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 62 毫秒
1.
3,4-Ethylene dioxythiophene (EDOT), as a monomer of commercial conductive poly(3,4-ethylene dioxythiophene) (PEDOT), has been facilely incorporated into a series of new π-conjugated polymer-based photocatalysts, i.e., BSO2–EDOT, DBT–EDOT, Py–EDOT and DFB–EDOT, through atom-economic C–H direct arylation polymerization (DArP). The photocatalytic hydrogen production (PHP) test shows that donor–acceptor (D–A)-type BSO2–EDOT renders the highest hydrogen evolution rate (HER) among the linear conjugated polymers (CPs) ever reported. A HER up to 0.95 mmol h−1/6 mg under visible light irradiation and an unprecedented apparent quantum yield of 13.6% at 550 nm are successfully achieved. Note that the photocatalytic activities of the C–H/C–Br coupling-derived EDOT-based CPs are superior to those of their counterparts derived from the classical C–Sn/C–Br Stille coupling, demonstrating that EDOT is a promising electron-rich building block which can be facilely integrated into CP-based photocatalysts. Systematic studies reveal that the enhanced water wettability by the integration of polar BSO2 with hydrophilic EDOT, the increased electron-donating ability by O–C p–π conjugation, the improved electron transfer by D–A architecture, broad light harvesting, and the nano-sized colloidal character in a H2O/NMP mixed solvent rendered BSO2–EDOT as one of the best CP photocatalysts toward PHP.

The excellent reactivity toward C–H direct arylation, water wettability and O–C p–π conjugation endow EDOT to be an attractive electron donor unit for CP photocatalysts, yielding an unprecedented hydrogen evolution rate up to 0.95 mmol h−1/6 mg catalyst.  相似文献   

2.
Correction for ‘Mass spectrometric detection of fleeting neutral intermediates generated in electrochemical reactions’ by Jilin Liu et al., Chem. Sci., 2021, 12, 9494–9499, DOI: 10.1039/D1SC01385H.

The authors regret that the details for ref. 15 and 17 were inadvertently swapped in the original article. The correct versions of ref. 15 and 17 are given below as ref. 1 and 2, respectively.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

3.
Correction for ‘Influence of the primary and secondary coordination spheres on nitric oxide adsorption and reactivity in cobalt(ii)–triazolate frameworks’ by Julia Oktawiec et al., Chem. Sci., 2021, DOI: 10.1039/d1sc03994f.

The authors regret that incorrect details were given for ref. 35, 37 and 59 in the original article. The correct versions of ref. 35, 37 and 59 are given below as ref. 1, 2 and 3, respectively.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

4.
Correction for ‘The oxygen-resistant [FeFe]-hydrogenase CbA5H harbors an unknown radical signal’ by Melanie Heghmanns et al., Chem. Sci., 2022, 13, 7289–7294, https://doi.org/10.1039/D2SC00385F.

The authors realized that incorrect references were cited following the sentence “In conjunction with the signal''s significant width, the frequency dependence clearly indicates spin–spin interaction between the F-clusters.” The correct references are shown below as ref. 1 and 2.Additionally ref. 36 and 37 were reversed in the reference list. The correct ref. 36 is shown below as ref. 3 and the correct ref. 37 is shown below as ref. 4.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

5.
Natural systems produce various γ-dicarbonyl-bearing compounds that can covalently modify lysine in protein targets via the classic Paal–Knorr reaction. Among them is a unique class of lipid-derived electrophiles – isoketals that exhibit high chemical reactivity and critical biological functions. However, their target selectivity and profiles in complex proteomes remain unknown. Here we report a Paal–Knorr agent, 4-oxonon-8-ynal (herein termed ONAyne), for surveying the reactivity and selectivity of the γ-dicarbonyl warhead in biological systems. Using an unbiased open-search strategy, we demonstrated the lysine specificity of ONAyne on a proteome-wide scale and characterized six probe-derived modifications, including the initial pyrrole adduct and its oxidative products (i.e., lactam and hydroxylactam adducts), an enlactam adduct from dehydration of hydroxylactam, and two chemotypes formed in the presence of endogenous formaldehyde (i.e., fulvene and aldehyde adducts). Furthermore, combined with quantitative chemoproteomics in a competitive format, ONAyne permitted global, in situ, and site-specific profiling of targeted lysine residues of two specific isomers of isoketals, levuglandin (LG) D2 and E2. The functional analyses reveal that LG-derived adduction drives inhibition of malate dehydrogenase MDH2 and exhibits a crosstalk with two epigenetic marks on histone H2B in macrophages. Our approach should be broadly useful for target profiling of bioactive γ-dicarbonyls in diverse biological contexts.

Natural systems produce various γ-dicarbonyl-bearing compounds that can covalently modify lysine in protein targets via the classic Paal–Knorr reaction.

Synthetic chemistry methods have been increasingly underscored by their potential to be repurposed as biocompatible methods for both chemical biology and drug discovery. The most-known examples of such a repurposing approach include the Staudinger ligation1 and the Huisgen-based click chemistry.2 Moreover, bioconjugation of cysteine and lysine can be built upon facile chemical processes,3 while chemoselective labelling of other polar residues (e.g., histidine,4 methionine,5 tyrosine,6 aspartic and glutamic acids7,8) requires more elaborate chemistry, thereby offering a powerful means to study the structure and function of proteins, even at a proteome-wide scale.The classical Paal–Knorr reaction has been reported for a single-step pyrrole synthesis in 1884.9,10 The reaction involves the condensation of γ-dicarbonyl with a primary amine under mild conditions (e.g., room temperature, mild acid) to give pyrrole through the intermediary hemiaminals followed by rapid dehydration of highly unstable pyrrolidine adducts (Fig. S1).Interestingly, we and others have recently demonstrated that the Paal–Knorr reaction can also readily take place in native biological systems.11–13 More importantly, the Paal–Knorr precursor γ-dicarbonyl resides on many endogenous metabolites and bioactive natural products.14 Among them of particular interest are isoketals15 (IsoKs, also known as γ-ketoaldehydes) which are a unique class of lipid derived electrophiles (LDEs) formed from lipid peroxidation (Fig. S2)16 that has emerged as an important mechanism for cells to regulate redox signalling and inflammatory responses,17 and drive ferroptosis,18 and this field has exponentially grown over the past few years. It has been well documented that the γ-dicarbonyl group of IsoKs can rapidly and predominantly react with lysine via the Paal–Knorr reaction to form a pyrrole adduct in vitro (Fig. 1).15 Further, the pyrrole formed by IsoKs can be easily oxidized to yield lactam and hydroxylactam products in the presence of molecular oxygen (Fig. 1). These rapid reactions are essentially irreversible. Hence, IsoKs react with protein approximately two orders of magnitude faster than the most-studied LDE 4-hydoxynonenal (4-HNE) that contains α,β-unsaturated carbonyl to generally adduct protein cysteines by Michael addition (Fig. S3).15 Due to this unique adduction chemistry and rapid reactivity, IsoKs exhibit intriguing biological activities, including inhibition of the nucleosome complex formation,19 high-density lipoprotein function,20 mitochondrial respiration and calcium homeostasis,21 as well as activation of hepatic stellate cells.22 Furthermore, increases in IsoK-protein adducts have been identified in many major diseases,23 such as atherosclerosis, Alzheimer''s disease, hypertension and so on.Open in a separate windowFig. 1The Paal–Knorr precursor γ-dicarbonyl reacts with the lysine residue on proteins to form diverse chemotypes via two pathways. The red arrow shows the oxidation pathway, while the blue one shows the formaldehyde pathway.Despite the chemical uniqueness, biological significance, and pathophysiological relevance of IsoKs, their residue selectivity and target profiles in complex proteomes remain unknown, hampering the studies of their mechanisms of action (MoAs). Pioneered by the Cravatt group, the competitive ABPP (activity-based protein profiling) has been the method of choice to analyse the molecular interactions between electrophiles (e.g., LDEs,24 oncometabolites,25 natural products,26,27 covalent ligands and drugs28–30) and nucleophilic amino acids across complex proteomes. In this regard, many residue-specific chemistry methods and probes have been developed for such studies. For example, several lysine-specific probes based on the activated ester warheads (e.g., sulfotetrafluorophenyl, STP;31N-hydroxysuccinimide, NHS32) have recently been developed to analyse electrophile–lysine interactions at a proteome-wide scale in human tumour cells, which provides rich resources of ligandable sites for covalent probes and potential therapeutics. Although these approaches can also be presumably leveraged to globally and site-specifically profile lysine-specific targets IsoKs, the reaction kinetics and target preference of activated ester-based probes likely differ from those of γ-dicarbonyls, possibly resulting in misinterpretation of ABPP competition results. Ideally, a lysine profiling probe used for a competitive ABPP analysis of IsoKs should therefore possess the same, or at least a similar, warhead moiety. Furthermore, due to the lack of reactive carbonyl groups on IsoK-derived protein adducts, several recently developed carbonyl-directed ligation probes for studying LDE-adductions are also not suitable for target profiling of IsoKs.33,34Towards this end, we sought to design a “clickable” γ-dicarbonyl probe for profiling lysine residues and, in combination with the competitive ABPP strategy, for analysing IsoK adductions in native proteomes. Considering that the diversity of various regio- and stereo- IsoK isomers15 (a total of 64, Fig. S2) in chemical reactivity and bioactivities is likely attributed to the substitution of γ-dicarbonyls at positions 2 and 3, the “clickable” alkyne handle needs to be rationally implemented onto the 4-methyl group in order to minimize the biases when competing with IsoKs in target engagement. Interestingly, we reasoned that 4-oxonon-8-ynal, a previously reported Paal–Knorr agent used as an intermediate for synthesizing fatty acid probes35 or oxa-tricyclic compounds,36 could be repurposed for the γ-dicarbonyl-directed ABPP application. With this chemical in hand (herein termed ONAyne, Fig. 2A), we first used western blotting to detect its utility in labelling proteins, allowing visualization of a dose-dependent labelling of the proteome in situ (Fig. S4). Next, we set up to incorporate this probe into a well-established chemoproteomic workflow for site-specific lysine profiling in situ (Fig. 2A). Specifically, intact cells were labelled with ONAyne in situ (200 μM, 2 h, 37 °C, a condition showing little cytotoxicity, Fig. S5), and the probe-labelled proteome was harvested and processed into tryptic peptides. The resulting probe-labelled peptides were conjugated with both light and heavy azido-UV-cleavable-biotin reagents (1 : 1) via CuI-catalyzed azide–alkyne cycloaddition reaction (CuAAC, also known as click chemistry). The biotinylated peptides were enriched with streptavidin beads and photoreleased for LC-MS/MS-based proteomics. The ONAyne-labelled peptides covalently conjugated with light and heavy tags would yield an isotopic signature. We considered only those modified peptide assignments whose MS1 data reflected a light/heavy ratio close to 1.0, thereby increasing the accuracy of these peptide identifications. Using this criterium, we applied a targeted database search to profile three expected probe-derived modifications (PDMs), including 13 pyrrole peptide adducts (Δ273.15), 77 lactam peptide adducts (Δ289.14), and 557 hydroxylactam peptide adducts (Δ305.14), comprising 585 lysine residues on 299 proteins (Fig. S6 and S7). Among them, the hydroxylactam adducts were present predominately, since the pyrrole formed by this probe, the same as IsoKs, can be easily oxidized when being exposed to O2. This finding was in accordance with a previous report where the pyrrole adducts formed by the reaction between IsoK and free lysine could not be detected, but rather their oxidized forms.37 Regardless, all three types of adducts were found in one lysine site of EF1A1 (K387, Fig. S8), further confirming the intrinsic relationship among those adductions in situ.Open in a separate windowFig. 2Adduct profile and proteome-wide selectivity of the γ-dicarbonyl probe ONAyne. (A) Chemical structure of ONAyne and schematic workflow for identifying ONAyne-adducted sites across the proteome. (B) Bar chart showing the distribution of six types of ONAyne-derived modifications formed in situ and in vitro (note: before probe labelling, small molecules in cell lysates were filtered out through desalting columns).State-of-the-art blind search can offer an opportunity to explore unexpected chemotypes (i.e., modifications) derived from a chemical probe and to unbiasedly assess its proteome-wide residue selectivity.38,39 We therefore sought to use one of such tools termed pChem38 to re-analyse the MS data (see Methods, ESI). Surprisingly, the pChem search identified three new and abundant PDMs (Fig. 1 and Table S1), which dramatically expand the ONAyne-profiled lysinome (2305 sites versus 585 sites). Overall, these newly identified PDMs accounted for 74.6% of all identifications (Fig. 2B and Table S2). Among them, the PDM of Δ287.13 (Fig. 1 and S7) might be an enlactam product via dehydration of the probe-derived hydroxylactam adduct. The other two might be explained by the plausible mechanism as follows (Fig. 1). The endogenous formaldehyde (FA, produced in substantial quantities in biological systems) reacts with the probe-derived pyrrole adduct via nucleophilic addition to form a carbinol intermediate, followed by rapid dehydration to a fulvene (Δ285.15, Fig. S7) and immediate oxidation to an aldehyde (Δ301.14, Fig. S7). In line with this mechanism, the amount of FA-derived PDMs was largely eliminated when the in vitro ONAyne labelling was performed in the FA-less cell lysates (Fig. 2B and Table S3). Undoubtedly, the detailed mechanisms underlying the formation of these unexpected PDMs require further investigation, and so does the reaction kinetics. Regardless, all main PDMs from ONAyne predominantly target the lysine residue with an average localization probability of 0.77, demonstrating their proteome-wide selectivity (Fig. S9).Next, we adapted an ABPP approach to globally and site-specifically quantify the reactivity of lysine towards the γ-dicarbonyl warhead through a dose-dependent labelling strategy (Fig. 3A) that has been proved to be successful for other lysine-specific probes (e.g., STP alkyne).31 Specifically, MDA-MB-231 cell lysates were treated with low versus high concentrations of ONAyne (1 mM versus 0.1 mM) for 1 h. Probe-labelled proteomes were digested into tryptic peptides that were then conjugated to isotopically labelled biotin tags via CuAAC for enrichment, identification and quantification. In principle, hyperreactive lysine would saturate labelling at the low probe concentration, whereas less reactive ones would show concentration-dependent increases in labelling. For fair comparison, the STP alkyne-based lysine profiling data were generated by using the same chemoproteomic workflow. Although 77.5% (3207) ONAyne-adducted lysine sites can also be profiled by STP alkyne-based analysis, the former indeed has its distinct target-profile with 930 lysine sites newly identified (Fig. S10 and Table S4). Interestingly, sequence motif analysis with pLogo40 revealed a significant difference in consensus motifs between ONAyne- and STP alkyne-targeting lysines (Fig. S11).Open in a separate windowFig. 3ONAyne-based quantitative reactivity profiling of proteomic lysines. (A) Schematic workflow for quantitative profiling of ONAyne–lysine reactions using the dose-dependent ABPP strategy (B) Box plots showing the distribution of R10:1 values quantified in ONAyne- and STP alkyne-based ABPP analyses, respectively. Red lines showing the median values. ***p ≤ 0.001 two-tailed Student''s t-test. (C) Representative extracted ion chromatograms (XICs) showing changes in the EF1A1 peptide bearing K273 that is adducted as indicated, with the profiles for light and heavy-labelled peptides in blue and red, respectively.Moreover, we quantified the ratio (R1 mM:0.1 mM) for a total of 2439 ONAyne-tagged lysines (on 922 proteins) and 17904 STP alkyne-tagged lysines (on 4447 proteins) across three biological replicates (Fig. S12 and Table S5). Strikingly, only 26.7% (651) of quantified sites exhibited nearly dose-dependent increases (R1 mM:0.1 mM > 5.0) in reactivity with ONAyne, an indicative of dose saturation (Fig. 3B and C). In contrast, such dose-dependent labelling events accounted for >69.1% of all quantified lysine sites in the STP alkyne-based ABPP analysis.31 This finding is in accordance with the extremely fast kinetics of reaction between lysine and γ-dicarbonyls (prone to saturation). Nonetheless, by applying 10-fold lower probe concentrations, overall 1628 (80.2%) detected lysines could be labelled in a fully concentration-dependent manner with the median R10:1 value of 8.1 (Fig. 3B, C, S12 and Table S5). Next, we asked whether the dose-depending quantitation data (100 μM versus 10 μM) can be harnessed to predict functionality. By retrieving the functional information for all quantified lysines from the UniProt Knowledgebase, we found that those hyper-reactive lysines could not be significantly over-represented with annotation (Fig. S12). Nonetheless, among all quantified lysines, 509 (25.1%) possess functional annotations, while merely 2.5% of the human lysinome can be annotated. Moreover, 381 (74.8%) ONAyne-labelled sites are known targets of various enzymatic post-translational modifications (PTMs), such as acetylation, succinylation, methylation and so on (Fig. S13). In contrast, all known PTM sites accounted for only 59.6% of the annotated human lysinome. These findings therefore highlight the intrinsic reactivity of ONAyne towards the ‘hot spots’ of endogenous lysine PTMs.The aforementioned results validate ONAyne as a fit-for-purpose lysine-specific chemoproteomic probe for competitive isoTOP-ABPP application of γ-dicarbonyl target profiling. Inspired by this, we next applied ONAyne-based chemoproteomics in an in situ competitive format (Fig. 4A) to globally profile lysine sites targeted by a mixture of levuglandin (LG) D2 and E2, two specific isomers of IsoKs that can be synthesized conveniently from prostaglandin H2 (ref. 41) (Fig. S2). Specifically, mouse macrophage RAW264.7 cells (a well-established model cell line to study LDE-induced inflammatory effects) were treated with 2 μM LGs or vehicle (DMSO) for 2 h, followed by ONAyne labelling for an additional 2 h. The probe-labelled proteomes were processed as mentioned above. For each lysine detected in this analysis, we calculated a control/treatment ratio (RC/T). Adduction of a lysine site by LGs would reduce its accessibility to the ONAyne probe, and thus a higher RC/T indicates increased adduction. In total, we quantified 2000 lysine sites on 834 proteins across five biological replicates. Among them, 102 (5.1%) sites exhibited decreases of reactivity towards LGs treatment (P < 0.05, Table S6), thereby being considered as potential targets of LGs. Notably, we found that different lysines on the same proteins showed varying sensitivity towards LGs (e.g., LGs targeted K3 of thioredoxin but not K8, K85 and K94, Table S6), an indicative of changes in reactivity, though we could not formally exclude the effects of changes in protein expression on the quantified competition ratios. Regardless, to the best of our knowledge, the proteome-wide identification of potential protein targets by IsoKs/LGs has not been possible until this work.Open in a separate windowFig. 4ONAyne-based in situ competitive ABPP uncovers functional targets of LGs in macrophages. (A) Schematic workflow for profiling LGs–lysine interactions using ONAyne-based in situ competitive ABPP. (B) Volcano plot showing the log2 values of the ratio between the control (heavy) and LGs-treated (light) channels and the −log10(P) of the statistical significance in a two-sample t-test for all quantified lysines. Potential targets of LGs are shown in blue (RC/T>1.2, P < 0.05), with the validated ones in red. (C) Bar chart showing the inhibitory effect of 2 μM LGs on the cellular enzymatic activity of MDH2. Data represent means ± standard deviation (n = 3). Statistical significance was calculated with two-tailed Student''s t-tests. (D) Pretreatment of LGs dose-dependently blocked ONAyne-labelling of MDH2 in RAW264.7 cells, as measured by western blotting-based ABPP. (E and F) LGs dose-dependently decreased the H2BK5 acetylation level in RAW 264.7 cells, as measured either by western blotting (E) or by immunofluorescence imaging (F). n = 3. For G, nuclei were visualized using DAPI (blue).We initially evaluated MDH2 (malate dehydrogenase, mitochondrial, also known as MDHM), an important metabolic enzyme that possesses four previously uncharacterized liganded lysine sites (K157, K239, K301 and K329, Fig. 4B) that are far from the active site (Fig. S14). We found that LGs dramatically reduced the catalytic activity of MDH2 in RAW264.7 cells (Fig. 4C), suggesting a potentially allosteric effect. We next turned our attention to the targeted sites residing on histone proteins, which happen to be modified by functionally important acetylation, including H2BK5ac (Fig. 4B) that can regulate both stemness and epithelial–mesenchymal transition of trophoblast stem cells.42 We therefore hypothesized that rapid adduction by LGs competes with the enzymatic formation of this epigenetic mark. Immunoblotting-based competitive ABPP confirmed that LGs dose-dependently blocked probe labelling of H2B (Fig. 4D). Further, both western blots and immunofluorescence assays revealed that LG treatment decreased the level of acetylation of H2BK5 (average RC/T = 1.3, P = 0.007) in a concentration-dependent manner (Fig. 4E and F). Likewise, a similar competitive crosstalk was observed between acetylation and LG-adduction on H2BK20 (average RC/T = 1.2, P = 0.01) that is required for chromatin assembly43 and/or gene regulation44 (Fig. 4B and S15). Notably, these findings, together with several previous reports by us and others about histone lysine ketoamide adduction by another important LDE, 4-oxo-2-noenal,11,45,46 highlight again the potentially important link between lipid peroxidation and epigenetic regulation. In addition to the targets validated as above, many other leads also merit functional studies considering diverse biological or physiologic effects of LGs in macrophages.  相似文献   

6.
Correction for ‘A photoexcited halogen-bonded EDA complex of the thiophenolate anion with iodobenzene for C(sp3)–H activation and thiolation’ by Tao Li et al., Chem. Sci., 2021, DOI: 10.1039/d1sc03667j.

Upon publication of the original article, the authors became aware of another work in this area that reported a related transformation, given below as ref. 1. This reference should also be considered when reading the original Chemical Science article.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

7.
Herein, we report for the first time single Au38 nanocluster reaction events of highly efficient electrochemiluminescence (ECL) with tri-n-propylamine radicals as a reductive co-reactant at the surface of an ultramicroelectrode (UME). The statistical analyses of individual reactions confirm stochastic single ones influenced by the applied potential.

Herein, we report for the first time single Au38 nanocluster reaction events of highly efficient electrochemiluminescence (ECL) with tri-n-propylamine radicals as a reductive co-reactant at the surface of a Pt ultramicroelectrode (UME).

Single entity measurements have been introduced by Bard and Wightman based on the collisions/reactions of individual nanoparticles and molecules at an ultramicroelectrode (UME).1–9 Since then, the field of single entity electrochemistry has gradually attracted several research groups and has become a frontier field of nanoelectrochemistry and electroanalytical chemistry.8,10–14 For instance, it has been shown that the chemistry of the electrode surface plays an important role in the collision/reaction events and the kinetics of reaction processes.15–21 Dasari et al. reported that hydrazine oxidation and proton reduction can be detected using single Pt nanoparticles on the surface of a mercury or bismuth modified Pt UME, and the material of the electrode was found to affect the shape of current–time transients.22,23 Fast scan cyclic voltammetry provides better chemical information about transient electrode–nanoparticle interactions, which is otherwise difficult to obtain with constant-potential techniques.24 There are only a few reports on photoelectrochemical systems including semiconductor nanoparticles designed to detect single nanoparticles in the course of photocatalysis processes.25–28 More importantly, owing to the nature of stochastic processes of single entity reactions, statistical analyses have shown substantial influences on the understanding of the underlying processes.Electrochemiluminescence or electrogenerated chemiluminescence (ECL),29 as a background-free technique,30–32 was also utilized to detect individual chemical reactions and single Pt nanoparticle collisions based on the reaction between the Ru(bpy)32+ complex and tri-n-propylamine (TPrA) radicals on the surface of an ITO electrode.2,33,34 It was found that the size of the nanoparticles, the origin of the interaction between particles and the electrode surface, the concentration of species generation, and the lifetime of individual electrogenerated nanocluster species (i.e., Au382+, Au383+, and Au384+) in conjunction with the reactivity of those oxidized species with co-reactant radical intermediates (i.e., TPrA radical) play crucial roles in the frequency of the ECL reaction events leading to individual ECL responses. More strikingly, a higher ECL reaction frequency is directly proportional to the amount of collected ECL light.21 Chen and co-workers also employed ECL to study stationary single gold-platinum nanoparticle reactivity on the surface of an ITO electrode.35 Lin and co-workers monitored the hydrogen evolution reaction in the course of “ON” and “OFF” ECL signals.36 Recently, we performed a systematic and mechanistic ECL study of a series of gold nanoclusters, with the general formula of Aun(SC2H4Ph)mz (n = 25, 38, 144, m = 18, 24, 60 and z = −1, 0, +1), where near-infrared (NIR) ECL emission was observed.37 There are several enhancement factors, such as catalytic loops38,39 that improve the signal to noise ratio. The Wightman group was able to report single ECL reactions based on the capability of ECL.7 Furthermore, thus far, we have explored ECL mechanisms and reported the ECL efficiency of five different gold nanoclusters i.e., Au25(SR)181−, Au25(SR)180, Au25(SR)181+, Au38(SR)240 and Au144(SR)600, among which the Au38(SR)240/TPrA system revealed outstanding ECL efficiency, ca. 3.5 times higher than that of Ru(bpy)32+/TPrA as a gold standard. Therefore, we decided to focus on the Au38 (SR)240/TPrA system. It was discovered that the ECL emission of these nanomaterials can be tuned through varying the applied potential and local concentration of the desired co-reactant.Herein, for the first time we report on ECL via a single Au38(SC2H4Ph)24 nanocluster (hereafter denoted as Au38 NC) reaction (eq. (1)) in the vicinity of an UME in the presence of TPrA radicals as a reductive co-reactant.1where x is the oxidation number that can be either 0, 1, 2, 3 or 4. Single ECL spikes (Fig. 1A) along with ECL spectroscopy were used for elucidating individual reaction events. Indeed, each single ECL spike demonstrates a single Au38(x−1)* reaction product. Au38 NCs were synthesized according to procedures reported by us and others, and fully characterized using UV-Visible-NIR, photoluminescence, 1HNMR spectroscopy and MALDI mass spectrometry to confirm the Au38 nanocluster synthesis (details are provided in ESI, Sections 1–3, Fig. S1–S4).38,40,41Fig. 2 (left) shows a differential pulse voltammogram (DPV) in an anodic scan of a 2 mm Pt disc electrode immersed in 0.1 mM Au38 acetonitrile/benzene solution containing 0.1 M TBAPF6 as the supporting electrolyte. There are five discrete electrochemical peaks at which Au380 was oxidized to Au38+ (E°′ = 0.39 V), Au382+ (E°′ = 0.60 V), and Au383+/4+ (E°′ = 0.99 V) and reduced to Au38 (E°′ = −0.76 V) and Au382− (E°′ = −1.01 V).38,40,41Open in a separate windowFig. 1(A) An example of the reaction event transient of 10 μM Au38 in benzene/acetonitrile (1 : 1) containing 0.1 M TBAPF6 in the presence of 20 mM TPrA at 0.9 V vs. SCE, acquired at 15 ms time intervals using a 10 μm Pt UME. The white dashed-line indicates the threshold to identify single ECL spikes. (B) Illustration of a single nanocluster ECL spike. (C) ECL instrumentation with an inset showing ECL spike generation in the vicinity of the Pt UME.Open in a separate windowFig. 2Anodic DPV for Au38 (left), reaction energy diagram of Au382+ and TPrA· (middle) along with the ECL–voltage curve (right) in an anodic potential scan at a 2 mm Pt disk electrode immersed in a solution of 10 μM Au38 with 20 mM TPrA.The rich electrochemistry of Au38 NCs is well-matched with that of co-reactants such as TPrA to generate near infrared-ECL (NIR-ECL), and the ECL emission efficiency of the Au38/TPrA system is 3.5 times larger than that of the Ru(bpy)32+/TPrA co-reactant ECL system.27Thus, it is of utmost interest to investigate the ECL generation of the above co-reactant system in single reactions, which improves the ECL signal detection sensitivity. To perform the ECL experiment a solution of 10 μM Au38 NC with 20 mM TPrA was prepared. We first confirmed the ECL light generation of such solution along with its blank solution containing only TPrA using a typical 2 mm diameter Pt disk electrode (Fig. 2, S5 and S6).A 10 μm Pt UME electrode, which is electrochemically inert (Fig. S7), was utilized to investigate the ECL of single NC reactions under potentiostatic conditions, at which a specific positive bias potential was applied to oxidize both Au38 and TPrA. Fig. 1A shows a typical ECL–time transient current curve (ECL intensity versus time) at 0.90 V vs. SCE, which was acquired using a photomultiplier tube (PMT, R928) for a duration of 1800 s at data acquisition time intervals of 15 ms (Fig. 1C and ESI, Section 3). Fig. 1B represents an exemplary event of a single ECL spike with a sharp increase followed by a decay in the ECL intensity. It is observed from the many spikes in Fig. 1B that this process can reoccur with a high probability in the vicinity of the UME, probably due to an electrocatalytic reaction loop (Fig. 1C). Indeed, ECL intensity was enhanced in this way as an already relaxed species, i.e., Au38z+1*, participates in an oxidation step to regenerate Au38z+1 to react with the TPrA radical (TPrA˙).Once photons resulting from the excited state relaxation in the vicinity of the UME are captured by the PMT, individual reaction events can be observed (Fig. 1A with the instrumentation schematic shown in Fig. 1C). As shown in Fig. 3A, there are many ECL spikes during 1800 s of measurement, each of which represents an individual ECL generation reaction in the vicinity of the UME surface. It is worth noting that there are several spikes with various intensities. This is most likely due to the Brownian motion which is random movement due to the diffusion of individual nanocluster species such as Au380, Au381+, Au382+, etc., electrogenerated at the local applied potentials. Long and co-workers42 proposed that silver nanoparticle collision on the surface of a gold electrode follows Brownian motion, leading to several types of surface-nanoparticle response peak shapes. In fact, the observed ECL spikes, shown in Fig. 1C, with a rise and an exponential decay suggested that Au38 nanocluster species diffuse directly through the electrode double-layer, move towards the tunneling region of the electrode surface, collide42 and become oxidized, react with TPrA radicals thereafter to produce excited states, and emit ECL. It is worth emphasizing that this path could be partially different for each individual nanocluster owing to the angle and direction relative to the electrode surface. The single Au38 NC reaction behaviour at various bias potentials was investigated following the electrochemical energy diagram shown in Fig. 2, middle. For example, at a bias potential of 0.70 V (the green spot on the DPV in Fig. 2), Au380 undergoes two successive oxidation reactions to Au382+ and TPrA oxidation and deprotonation start to generate TPrA·. In fact, at a very close oxidation potential to Au382+, TPrA is also oxidized to its corresponding cation radical (ca. 0.80 V vs. SCE) Fig. S6, followed by deprotonation to form the TPrA radical.38 The TPrA· with a very high reduction power (E°′ = −1.7 eV)43 injects one electron to the LUMO orbital of the nanocluster and forms excited state Au38+*, as illustrated in the reaction energy diagram in Fig. 2, middle.38 Then, Au38+* emits ECL light while relaxing to the ground state. For another instance, at 1.10 V vs. SCE (the red spot on the DPV in Fig. 2), Au380 is oxidized to Au383/4+ feasibly. At this potential, the TPrA radical is generated massively in the vicinity of the electrode. The efficient electron transfer between the TPrA radical and Au383/4+ generates both Au382+* and Au383+* that emit light at the same wavelength of 930 nm.38 The results of such interactions produced a transient composed of many ECL events (Fig. 3A), which is an indication of bias potential enforcement on the nanocluster light emission.Open in a separate windowFig. 3Single-nanocluster ECL photoelectron spectroscopy of Au38. ECL–time transients (A), statistics of the number of photons (B), histogram of the single reaction time between sequential spikes (C) and accumulated ECL spectrum (D) for a 10 μm Pt UME at 1.1 V immersed in a 10 μM Au38 nanocluster solution in benzene/acetonitrile (1 : 1) containing 0.1 M TBAPF6 in the presence of 20 mM TPrA. (E)–(H) The counterpart plots to (A)–(D) for the UME biased at 0.7 V. # represents the number.We further tried to collect the current–time traces of such events; however, owing to the high background current originating from the high concentration of TPrA relative to that of the nanocluster, no noticeable spikes in the current were observed.In order to study the photochemistry and understand deeply the single nanocluster reactions, ECL–time transients were collected at different applied potentials (i.e., 0.7, 0.8, 0.9 and 1.1 V vs. SCE) as labelled in green, brown, purple, and red on the DPV in Fig. 2, respectively. The transients were further analysed using our home-written MATLAB algorithm adapted from that for nanopore electrochemistry.44 The population of individual events was identified by applying an appropriate threshold to discriminate ECL spikes from the noise as demonstrated in Fig. 1A. In fact, the applied algorithm also assisted us to learn about the raising time and intensity of each spike, as well as photons of individual spikes. For instance, Fig. 3A shows another typical transit for 1800 s at an UME potential bias of 1.1 V for the ECL events. Indeed, the integrated area of each peak, the charge of the photoelectrons at the PMT, is directly proportional to the number of photons emitted from individual reactions (see ESI, Section 5). Basically, the PMT amplifies the collected single photon emitted in the course of light-to-photoelectron conversion (see ESI, Section 6 and Fig. S8) and translates a single photon into photoelectrons. The extracted charge of each ECL reaction, QECL, was then converted to the corresponding number of photons by dividing by the gain factor, g, which is 1.55 × 106 (Fig. S8), following eqn (2):2The histograms of the number of photons show a Gaussian distribution (Fig. 3B) with a reaction frequency of 53.5 ± 2.9 at E = 1.1 V, whereas at a lower potential of 0.7 V the reaction frequency drops to 18.5 ± 1.7 (Fig. 3F). This indicates that there is a three-fold lower reaction occurrence at the lower potential. The integration of the Gaussian fitting at 1.1 V and 0.7 V also reveals a three-fold drop from 3.3 × 105 to 1.2 × 105 photons over 1800 s.To further explore the effect of electrode potential bias on the single Au38 NCs ECL reaction, potentials lower than 1.1 and higher than 0.7 V, ca. 0.8 and 0.9 V (brown and purple labels in Fig. 2), were applied. In fact, the resulting ECL–time transients show a lower population of single spikes (Fig. S12A and ESI,). The integrated Gaussian curve values support the ECL–time transient observations with ∼4.1 × 104 and ∼6.5 × 104 photons, respectively. In fact, it is unlikely that the PMT would get more than two events in the duration, owing to the following reasons: (i) it has been shown that only 5.5% of incoming photons can be effectively converted to photoelectron signals by our R928 PMT during our absolute efficiency calibration, ESI Section 6 and Fig. S8–S19;45 (ii) spherical ECL emission is proven to be detected for a substantial small part upon examination of our detection system for the absolute ECL quantum efficiency;45 (iii) Au38 nanocluster ECL emissions occur at 930 nm, which is almost at the wavelength detection limit of our PMT response curve.38,45In addition, we evaluated the stochastic (a series of random events at various probability distributions) nature of the observed events and extracted the reaction time interval (τ) at various potentials. The resulting graph shows an exponential decay (Fig. 3C) as expressed in eqn (3):3where frequency (λ) gives the mean rate of the event and A represents the fitting amplitude. One can expect to obtain the distribution of the number of emitted photons and spatial brightness function. In fact, the exponential decay is a clear indication of random single reaction events as Whiteman and co-workers described for a 9,10-diphenylanthracene (DPA) ECL system in the annihilation pathway.7,46 At a potential of 1.1 V, λ and A are found to be 4.98 ± 0.02 ms−1 and 80.4 ± 3.2, whereas at 0.7 V, λ and A turned out to be 32.9 ± 1.6 ms−1 and 9.5 ± 0.1 (Fig. 3C and G). Indeed, the lower potential of 0.70 V vs. SCE is high enough to generate the TPrA radical along with Au382+, thereby leading to excited Au38+*, Fig. 3E. One can conclude that at the applied potentials of 0.7 V and 1.1 V, Au380 is oxidized to Au382+ and Au384+, resulting in the generation of Au38+* and Au383+* under static conditions. Thus, there are higher populations of ECL spikes with no discrepancy in the number of collected photon distributions. However, at two intermediate potentials, i.e., 0.8 and 0.9 V, a dynamic behaviour which is due to the mixed oxidation of Au38 species, in the vicinity of the UME, is observed. In fact, at these two applied potentials, the local concentration of the corresponding gold nanoclusters (i.e., Au383+ and Au384+) is not sufficient to produce significant ECL spikes. We also attempted to collect the ECL spectrum using a charge-coupled device (CCD) camera, which is relatively more sensitive in the NIR region (e.g., λ > 900 nm, Fig. S16). Fig. 3D and H display an accumulated spectrum at 1.1 and 0.7 V vs. SCE, which is collected for 30 minutes. The fitted accumulated ECL spectrum indicates an ECL peak emission at 930 nm and supports higher reactivity at 1.1 V than that at 0.7 V.38 To confirm that the observed ECL spikes and accumulated spectra are generated based on the oxidation of Au38 nanoclusters in the presence of TPrA radicals, ECL–time transients were recorded upon holding an applied potential at which no faradaic process occurs. Fig. S11 represents ECL–time curves and accumulated ECL spectra at 0.0 V and 0.4 V. One can notice that no appreciable ECL signal can be observed.In addition, we investigate the Pearson cross-correlation (ρ) between the intensities of ECL spikes with τ as shown in Fig. S14 in which there is a positive correlation at 0.7 and 1.0 V and a negative correlation at 0.8 and 0.9 V. In fact, ρ evaluates whether there is a stationary random process between the two defined parameters (see ESI, Section 6). Interestingly, the frequency of the reaction at different applied potentials revealed decay from 0.7 to 0.8 V, followed by an upward trend to 0.9 and 1.1 V vs. SCE (Fig. S15). This could be additional support for the transition stage at 0.8 and 0.9 V, where the applied potential as the major driving force to generate oxidized forms (e.g., Au383+ and Au384+) governs the flux of the nanocluster species that reach the vicinity of the electrode. Furthermore, the effectiveness of electron transfer reaction kinetics between the radical species, i.e., Au38z+1 and TPrA radical, competes with the flow of the incoming nanoclusters. It is worth mentioning that each of the ECL single event experiments was repeated three times, and very similar results were obtained. Moreover, lower (5 μM) and higher (20 μM) concentrations of Au38 in the presence of 20 mM were tested. In fact, the former shows a smaller number of single reactions; however the later revealed a larger number of multiple reactions (Fig. S13).In summary, in this communication we demonstrated that Au38 NC ECL at the single reaction level can be monitored using a simple photoelectrochemical setup following a straightforward protocol. Indeed, we have rich basic knowledge about the ECL mechanisms of various gold nanoclusters with different charge states (Au25(SR)181+, Au25(SR)180, Au25(SR)181−) and various sizes (Au25(SR)180, Au38(SR)240, Au144(SR)600) in fine detail. Thus, the ECL emission mechanisms of gold clusters, including the contribution of each charge state and influence of various concentrations of co-reactants, are well known. For instance, in our previous studies38,39,47–49 we clearly identified three charge states of an Au25(SR)181−/TPrA system and we discovered that at a high concentration of TPrA the reduction in the bulk solution of gold nanoclusters influences the ECL emission wavelength. We also have learnt that the Au38/TPrA system is a co-reactant independent of co-reactant concentration. Furthermore, an extensively higher concentration of TPrA provides a dominant reaction over any unknown decomposition reaction at higher oxidation states of Au38. It was discovered that the population of ECL reactions is directly governed by the applied bias potential on a Pt UME. This work is a strong indication of the high sensitivity of the ECL technique in detecting single ECL reactions in a simple solution, which complements those reported by the Bard group using rubrene, for instance, embedded in an organic emulsion in the presence of TPrA or oxalate as a co-reactant.50,51 These systems needed a substantial ECL enhancement in the presence of an ionic liquid as the supporting electrolyte and emulsifier. The current approach can be further extended to investigate other molecules and nanomaterials'' electrocatalytic processes at the single reaction level.  相似文献   

8.
9.
The redox chemistries of both the bromide oxidation and bromine reduction reactions are studied at single multi-walled carbon nanotubes (MWCNTs) as a function of their electrical potential allowing inference of the electron transfer kinetics of the Br2/Br redox couple, widely used in batteries. The nanotubes are shown to be mildly catalytic compared to a glassy carbon surface but much less as inferred from conventional voltammetry on porous ensembles of MWCNTs where the mixed transport regime masks the true catalytic response.

Schematic of a carbon nanotube impact in bromide solution.

The bromine–bromide redox couple plays an essential role in diverse energy storage devices including hydrogen–bromine, zinc–bromine, quinone–bromine, vanadium–bromide and bromide–polysulphide flow batteries.1–5 The Br2/Br redox couple is attractive as a cathode reaction due to its high standard potential, large solubility of both reagents, high power density and cost efficiency.6 The performance of such devices is generically limited by the thermodynamics and kinetics of the redox couple comprising the battery with fast (‘reversible’) electron transfer is essential. In many cases, including the Br2/Br couple the electrode reaction involves more than one electron as given in the stoichiometric reaction:2Br − 2e ⇄ Br2; E0 = 1.08 V vs. SHEwith, at high bromide concentrations, the possibility of the follow up chemical reaction7Br2 + Br ⇄ Br3Since electrons are usually transferred sequentially this implies that the mechanism is multistep with any of the individual mechanistic steps in principle being rate limiting. For this reason catalysts are commonly required to enhance the electrode kinetics at otherwise favourable electrode materials. One type of catalyst which has seen wide usage, including for the Br2/Br couple8,9 are carbon nanotubes (CNTs) with suggested advantages which include high surface area and the inherent porosity of CNT composites.10 The deployment of CNTs as a porous composite presents a further level of complexity to the electrode reaction beyond its multistep character because of the ill-defined mass transport within the porous layer. In particular ascertaining the intrinsic electron transfer kinetics and hence the level of catalysis, if any, is essentially impossible since these are masked in the voltammetric response by diffusional mass transport effects.11–14 Specifically the transport within the porous structure of CNT layers is dominated by thin-layer and other15,16 effects which give the illusion of electrochemical reversibility. In order to unscramble possible electro-catalysis of the bromine/bromide couple a different approach is needed.In the following we study both the electro-oxidation of bromide (BOR) and the electro-reduction of bromine (BRR) at single MWCNTs via ‘nano-impact (aka ‘single entity’) electrochemistry’17–20 in aqueous solution. In this approach a micro-wire electrode at a fixed potential is inserted in a suspension of CNTs in the solution of interest. From time to time a single CNT impacts the electrode, adopts the potential of the latter for the duration of the impact which in the case of CNTs can vary from 1–100 of seconds21–23 and sustained catalytic currents flow if the oxidation/reduction of interest is faster at the nanotube in comparison with the micro-wire electrode. The catalytic currents are studied as a function of potential revealing the electron transfer kinetics. Fig. 1 shows the concept of the experiment.Open in a separate windowFig. 1Schematic representation of ‘nano-impact’ electrochemistry on a carbon micro wire electrode for the oxidation of aqueous bromide from which the kinetics of the BOR are inferred. Analogous experiments but showing negative impact currents allow the inference of the kinetics of the BRR.The BOR and BRR were studied first, however, voltammetrically at an unmodified glassy carbon (GC) electrode as shown in Fig. 2 (black line) using 5.0 mM solutions of either NaBr or Br2 in 0.1 M HNO3. The midpoint potential was 0.82 V versus the saturated calomel electrode (SCE) consistent with the literature values for the formal potential of the Br2/Br couple.24 The voltammograms were analysed to give transfer coefficients of 0.45 ± 0.01 and 0.33 ± 0.01 (ESI, Section 2) for the BOR and BRR respectively. Both processes were inferred to be diffusional and the diffusion coefficients DBr and DBr2 were calculated to be 2.05 (±0.04) × 10−5 cm2 s−1 and 1.50 (±0.04) × 10−5 cm2 s−1 (ESI, Section 3) using the Randles–Ševčík equation for an irreversible reaction the values are consistent with literature reports.24 Then the electrodes were modified with 30 μg of MWCNTs consisting of ca. 125 monolayers (the calculation is given in the ESI, Section 9) of MWCNTs assuming that they are closely packed across the area of the GC electrode, and the resulting voltammograms are shown in Fig. 2 (red line). In comparison with the unmodified electrode, enhanced currents are seen for the Br2/Br couple which partly reflects the enhanced capacitance of the interface reflecting in turn the large surface area of the deposited nanotubes (ca. 60–120 cm2). Larger signals are also seen indicating a thin layer contribution from the material occluded within the porous layer which also leads to the apparently quasi-reversible shape of the voltammograms obtained for both reactions. A log–log plot of peak current (Ip) vs. scan rate (ν) showed a gradient value of 0.68 (±0.01) and 0.66 (±0.03) for the BOR and BRR (ESI, Section 4) confirming a mixed mass transport regime12,14 with a combination of semi-infinite diffusion and thin layer behaviour. The transition from the fully irreversible to the apparent quasi-reversible character is sometimes confused with electro-catalysis attributed to the CNTs rather than thin-layer diffusion. In order to ascertain the true catalytic response, single entity electrochemistry was measured to obtain the BOR and BRR responses at single CNTs.Open in a separate windowFig. 2Cyclic voltammograms at pristine GC (black line) and 30 μg MWCNTs dropcast on GC (red line) at a scan rate of 0.05 V s−1 (a) for the bromide oxidation reaction (BOR) in 5.0 mM NaBr in 0.1 M HNO3, (b) for the bromine reduction reaction (BRR) in 5.0 mM bromine in 0.1 M HNO3.For single entity measurements, a clean carbon wire (CWE, length 1 mm and diameter 7 μm) working electrode was used. Chronoamperograms were recorded at a constant applied potential of 0.2 V vs. SCE and 1.3 V vs. SCE for the BOR and BRR respectively (5.0 mM solutions). These values were selected in the light of Fig. 2 to provide a large overpotential for each reaction. Clear oxidative and reductive current steps were observed (Fig. 3). These were ascribed to the arrival of a MWCNT at the electrode surface and the resulting catalytic electron transfer for the duration of the impact. No steps were observed in the absence of MWCNTs (ESI, Fig. S4). The average residence time of the MWCNT was 1.2 (±0.5) seconds and the frequency of the collisions was 0.3 (±0.1) impacts per second. The average impact current for the BOR at 1.3 V vs. SCE was 2.8 (±0.2) nA (65 impacts) and for the BRR at 0.2 V vs. SCE it was 3.8 (±0.1) nA (70 impacts). The impact currents were assumed to be entirely faradaic since control experiments in 0.1 M HNO3 solution in the presence of 100 μg of MWCNTs (in the absence of Br and Br2) showed no obvious impacts as shown in ESI Section 10.Open in a separate windowFig. 3Chronoamperograms showing the impact step current (a) for the BOR in 5.0 mM NaBr in 0.1 M HNO3 at 1.3 V vs. SCE, (b) for the BRR in 5.0 mM bromine in 0.1 M HNO3 at 0.2 V vs. SCE.Further, impacts for both the BOR and BRR were observed at various potentials (ESI, Section 11) and analysed to obtain the average faradaic current at each potential. The average impact step current was plotted against the applied potential (Fig. 4). Two sigmoidal curves were obtained reflecting the current–potential response for either the bromide oxidation (BOR) or the bromine reduction (BRR). The curves reflect the average voltammograms (current–potential characteristics) for the Br2/Br redox reaction at single carbon nanotubes. The shape of the two sigmoidal curves reflects the onset of electrolysis followed by a diffusion controlled plateau at high over-potentials.25 Mass transport corrected Tafel analysis (Fig. 4; inset) showed the transfer coefficients β to be ca. 0.42 and α to be ca. 0.20 from the impacts for the BOR and BRR respectively (ESI, Section 6). The length distribution of the MWCNTs was calculated (ESI, Section 6) from the currents recorded at potentials corresponding to the plateau in Fig. 4 assuming that the reactions are (Fickian) diffusion controlled at the potentials used and by modelling the CNTs as cylindrical electrodes21 assuming a nanotube radius of 15 (±5) nm and the diffusion coefficients reported above. Chronoamperometry was also conducted for the BOR and BRR in the absence of MWCNTs at 1.3 V and 0.2 V vs. SCE respectively to confirm that no impact currents were contributed by the redox species in the electrolyte (ESI, Section 5). Alongside, chronoamperograms in 0.1 M HNO3 and 100 μg show that the impact current was contributed only by the Br and Br2 redox reaction and the results are shown in the ESI, Section 10.Open in a separate windowFig. 4Average step currents observed as a function of applied potential (a) for the BOR in 5.0 mM NaBr in 0.1 M HNO3 at, (b) for the BRR in 5.0 mM Bromine in 0.1 M HNO3; insets in both the cases show mass transport corrected Tafel analyses.The lengths were found to be 5.4 (±3.4) μm (BOR) and 5.9 (±1.3) μm (BRR) and are given in Fig. 5 (see ESI, Section 7 for calculations). These values were compared with previously reported dark-field optical microscopy data and good agreement was observed with the literature value of 5.3 (±2.1) μm.26 The observed consistency provides strong support for the choice of modelling the single entity voltammetry by analogy with that of a cylindrical electrode.Open in a separate windowFig. 5The length of MWCNTs calculated from the impact currents for the BOR (at 1.3 V vs. SCE) and BRR (at 0.2 V vs. SCE).It is evident that the single entity measurements allow a clear analysis of the catalytic behaviour of the carbon nanotubes by providing a well-defined diffusional regime conducive to the extraction of the electrode kinetics of both the bromide oxidation and the bromine reduction process. In contrast, electrodes were formed by ensembles of carbon nanotubes in the form of a porous layer where the mixed transport regime is not amenable to ready modelling and the dissection of thin-layer effects from the measured voltammetry. The electron transfer kinetics for both the BOR and BRR at single MWCNTs was then obtained via full simulation of the two single entity ‘voltammograms’ using the above measured diffusion coefficients and again treating the impacted MWCNT as a cylindrical electrode with uniform diffusional access and further assuming Butler–Volmer kinetics. For the BOR, one electron transfer was considered as given below,For the BRR the two electron transfer was modelled as,Br2 + 2e → 2BrThe set of parameters used for the analysis are given in the ESI, Section 8. By using the transfer coefficients deduced from Fig. 4, the only unknown is the standard electrochemical rate constant k which is determined by fitting the impact voltammogram measured relative to a formal potential for the Br2/Br couple of 0.82 V vs. SCE obtained from the voltammogram at pristine GC. Fig. 6 shows the fitting for the BOR and the BRR with rate constants kBOR of 1.0 (±0.1) × 10−3 cm s−1 and kBRR of 5.0 (±0.1) × 10−4 cm s−1 respectively. The transfer coefficients and rate constants obtained from impacts were compared to the voltammograms obtained at pristine GC for the BOR and BRR and are given in Open in a separate windowFig. 6DIGISIM simulated curves (black line) for average impact currents obtained at different potentials (red circles) (a) for the BOR with a rate constant (kBOR) of 1.0 (±0.1) × 10−3 cm s−1; (b) BRR with a kBRR of 5.0 (±0.1) × 10−4 cm s−1.Transfer coefficients and rate constants for the BOR in 5.0 mM NaBr in 0.1 M HNO3 and the BRR in 5.0 mM bromine in 0.1 M HNO3 obtained at the glassy carbon macroelectrode GC, and single MWCNT impact current
Analysed parameterOxidation of bromideReduction of bromine
Transfer coefficient (GC)β = 0.45α = 0.33
Transfer coefficient (impact current)β = 0.42α = 0.20
kBOR/cm s−1 (GC)9.5 (±0.1) × 10−52.0 (±0.1) × 10−5
kBRR/cm s−1 (impact current)1.0 (±0.1) × 10−35.0 (±0.1) × 10−4
Open in a separate windowIn summary, MWCNTs were studied for their catalytic behaviour towards the Br2/Br redox couple. From the drop-cast experiment, the ensemble of MWCNTs showed mixed mass transport behaviour complicating and precluding the elucidation of their catalytic behaviour. In contrast, single nano-impact electrochemistry of MWCNTs shows faster electrochemical rate constants compared to pristine GC. This confirms the catalytic activity of MWCNTs for the Br2/Br redox reaction but the values determined are insufficiently enhanced over glassy carbon leaving considerable room for improvement via the use of alternative electrocatalysts to carbon nanotubes.  相似文献   

10.
11.
12.
Expediting the oxygen evolution reaction (OER) is the key to achieving efficient photocatalytic overall water splitting. Herein, single-atom Co–OH modified polymeric carbon nitride (Co-PCN) was synthesized with single-atom loading increased by ∼37 times with the assistance of ball milling that formed ultrathin nanosheets. The single-atom Co-N4OH structure was confirmed experimentally and theoretically and was verified to enhance optical absorption and charge separation and work as the active site for the OER. Co-PCN exhibits the highest OER rate of 37.3 μmol h−1 under visible light irradiation, ∼28-fold higher than that of common PCN/CoOx, with the highest apparent quantum yields reaching 4.69, 2.06, and 0.46% at 400, 420, and 500 nm, respectively, and is among the best OER photocatalysts reported so far. This work provides an effective way to synthesize efficient OER photocatalysts.

Single-atom CoII-OH modified polymeric carbon nitride synthesized with increased single-atom loading under the assistance of ball milling exhibits high photocatalytic water oxidation activity with Co-N4OH as the highly active site.

Massive fuel energy consumption induced environmental and ecological problems, especially the greenhouse effect, and the resultant extreme climates and rise in sea level are threatening human life.1 As a potential substitution for fuel energy, hydrogen energy conversion from solar energy via photocatalytic water splitting attracts great attention from scientists.2–5 However, the photocatalytic hydrogen evolution efficiency from overall water splitting is still restricted by the sluggish oxygen evolution reaction (OER) that involves energy absorption, four-electron transfer, breakage of O–H bonds, and formation of O–O bonds,6,7 and thus efficient OER photocatalysts become the key to achieving efficient overall water splitting. Though numerous hydrogen evolution photocatalysts have been reported, research on OER photocatalysts is mainly around a few semiconductors including BiVO4, WO3, Ag3PO4, α-Fe2O3, etc.8–11 and their activity is not high enough yet for practical applications. Therefore, exploring high-efficiency OER photocatalysts is still necessary.Polymeric carbon nitride (PCN) was first reported in 2009 (ref. 12) as a photocatalyst with a layered melon-type orthorhombic structure,13 and thereafter quickly became a “star” photocatalyst thanks to its advantages of being visible-light responsive and metal-free, non-toxic, and low cost, and its relatively high chemical stability.14 Because of several self-deficiencies including fast photogenerated charge recombination and a narrow optical absorption spectrum, PCN exhibits relatively low photocatalytic activity.15 Then, a series of strategies were put forward successively to enhance the photoactivity of PCN, such as enhancement of crystallinity,16 morphological control,17 structural modification18 (including extensively researched single atom modification in recent years19,20), exfoliation,21 construction of hetero-(homo-)junctions,22 and loading of noble metals.23 Though photocatalytic water splitting on PCN was extensively researched in the past, the research was mainly around the hydrogen evolution half-reaction used for exploring properties and the catalytic mechanism of photocatalysts, and little research was focused on the industrially useable overall water splitting process owing to the sluggish OER.15 Therefore, enhancing the photocatalytic OER activity of PCN becomes the key to practical applications.To increase OER rates of PCN, several kinds of methods were proposed, such as rational design of compound cocatalysts (e.g., CoOx, IrO2, CoP, CoPi, RhOx, RuOx, PtOx, MnOx, Co(OH)2, Ni(OH)2, and CoAl2O4 (ref. 24–30)), modification of carbon dots and carbon rings,31,32 fabrication of special architectures of PCN (e.g., PCN quantum dot stacked nanowires33), and single-atom (e.g., B, Co, and Mn34–36) modification. For instance, Zhao and coauthors prepared B and N-vacancy comodified PCN that exhibits the highest OER rate of ∼28 μmol h−1 (ref. 36) and recently their group further used these B doped PCN ultrathin nanosheets to fabricate a Z-scheme heterojunction for overall water splitting with a solar-to-hydrogen efficiency reaching ∼1.2%.37 Comparatively, PCN loaded with compound cocatalysts can only enhance OER activity to a limited degree and there are finite methods for carbon modification and special architecture fabrication. Single-atom modification shows a bright prospect, on account of metal atoms capable of being inserted into the framework of PCN and effectively increasing the OER activity. However, reported single metal atom modification routes are all based on direct ion adsorption on PCN or calcination of mixtures of metal salts and PCN feedstocks.34,35,38 New routes need be explored to increase effective loading of single atoms in PCN. Besides, the metal-OH structure is considered efficient for the OER,30,39,40 and a single metal atom-OH structure has never been reported for modification of PCN, though Mn–OH was thought to play a key role in the OER process.34Ball milling is an extensively used versatile and scalable way for preparation of heterogeneous catalysts and even single-atom catalysts,41,42 but was rarely used in synthesis of PCN-based single-atom photocatalysts. In this work, we synthesized single-atom Co–OH modified PCN (Co-PCN) with the single-atom content in PCN highly increased with the assistance of ball milling. The simple synthetic route is shown in Fig. 1a. PCN was ball-milled to obtain BM-PCN that then adsorbed Co2+ till saturation to form BM-PCN/Co which was calcined to obtain BM-PCN/Co-c (Co-PCN). For comparison, PCN was directly used to adsorb Co2+ till saturation to form PCN/Co which was calcined to obtain PCN/Co-c. PCN mainly comprises large blocks with the size of several micrometers (Fig. S1), while BM-PCN contains massive irregular particles with the size reduced to several hundreds of nanometers (Fig. S2), indicative of high efficacy of ball milling. BM-PCN/Co-c exhibits a similar morphology as BM-PCN (Fig. 1b and S3) and PCN/Co-c exhibits a similar morphology to PCN (Fig. S4), but the surface area and mesopore volume of BM-PCN and BM-PCN/Co-c are not higher than those of PCN and PCN/Co-c (Fig. S5), manifesting that ball-milling and subsequent calcination did not form massive mesopores, which accords well with the particle size variation from several micrometers (before ball milling) to several hundreds of nanometers (after ball milling). However, the Co content in BM-PCN/Co-c, BM-PCN/Co, PCN/Co-c, and PCN/Co was measured to be 0.75, 0.50, 0.02, and ∼0.02 wt%, respectively, by inductively coupled plasma mass spectrometry (ICP-MS). The ∼37 times higher Co content in BM-PCN/Co-c than in PCN/Co-c suggests the ball-milling enhanced adsorption of Co2+ on surfaces of BM-PCN, which should arise mainly from the ball-milling induced increase of surface energy and adsorption sites.43Open in a separate windowFig. 1(a) Schematic illustration for synthesis of single-atom CoII-OH modified PCN (BM-PCN/Co-c); and (b) SEM, (inset in b) TEM, (c) AFM, (d) EDS elemental mapping, and (e) HAADF-STEM images of BM-PCN/Co-c.The TEM image shows the existence of small and ultrathin nanosheets in BM-PCN/Co-c (inset in Fig. 1b) which can also be observed in the atomic force microscopy (AFM) image with a thickness of ∼7–10 nm and lateral size of <70 nm (Fig. 1c), and formation of these ultrathin nanosheets results from the ball milling of PCN.44 It should be noted that most formed ultrathin nanosheets with high surface energy may stack into compact particles upon ball milling, leading to no increase of the total surface area. Energy dispersive X-ray spectroscopy (EDS) elemental mapping images of BM-PCN/Co-c indicate homogeneous distribution of C, N, O, and Co elements in the sample (Fig. 1d). The high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image of BM-PCN/Co-c shows massive white spots (marked by circles) with a mean size of <1 Å dispersed in the sample (Fig. 1e and S6), which should correspond to single-atom Co.To further verify the single-atom Co structure in BM-PCN/Co-c, Co K-edge X-ray absorption near-edge structure spectroscopy (XANES) and extended X-ray absorption fine structure (EXAFS) analysis were performed. As shown in Fig. 2a, the absorption-edge position of BM-PCN/Co-c is quite close to that of CoO and their peak positions are similar and far from those of other reference samples, indicating that the valence of Co in BM-PCN/Co-c is about +2. The bonding structure around Co was determined by Fourier transformed (FT) k3-weighted EXAFS analysis. As shown in Fig. 2b, a distinct single Co-ligand peak at ∼1.6 Å for BM-PCN/Co-c is observed, which prominently differs from the Co–Co coordination peak at ∼2.2 Å for Co foil and the CoII–O coordination peak at ∼1.7 Å for CoO. The wavelet transform (WT) contour plot of BM-PCN/Co-c shows only one intensity maximum (Fig. S7), and the Cl 2p core-level XPS spectrum of BM-PCN/Co-c reveals no residue of Cl (Fig. S8). These further indicate the single-atom dispersion of Co species. Apparently, the Co-ligand peak is almost consistent with the CoII–N peak for Co porphyrin, suggesting that the single-atom Co in BM-PCN/Co-c mainly coordinates with N. Least-square EXAFS curve fitting was performed to confirm quantitative structural parameters of CoII in BM-PCN/Co-c, as shown in Fig. 2c, S9, and S10 and Table S1. Simple Co–N single-shell fitting of BM-PCN/Co-c (Fig. S10) gave a coordination number of 5.6 ± 0.4 (Table S1), that is, CoII coordinates with five atoms. Considering that the PCN monolayer provides four appropriate N coordination sites at most,45 CoII likely coordinates with four N atoms and one OH atom. Thus, we further performed Co–N4/Co–O double-shell fitting (Fig. 2c) and the obtained R-factor (0.0011) remarkably reduces relative to that from Co–N single-shell fitting (0.0035), indicative of rationality of the proposed CoII–N4OH structure. Confirmed Co–N and Co–O bond lengths are 2.04 and 2.15 Å, respectively (Table S1).Open in a separate windowFig. 2(a) Co K-edge XANES and (b) EXAFS spectra of Co foil, Co porphyrin (Copr), CoO, Co3O4, Co2O3, and BM-PCN/Co-c; EXAFS (c) R space-fitting and (inset in c) K space-fitting curves of BM-PCN/Co-c; (d) optimized structure of PCN and Co-doped PCN with different doping configurations and calculated formation energies (e) of Co doped PCN; and (e) Co 2p and (f) O 1s core-level XPS spectra of samples.To further confirm rationality of the Co–N4OH coordination structure, density functional theory (DFT) calculations were conducted. As shown in Fig. 2d, three possible CoII coordination structures in the PCN monolayer were explored. The Co–N4OH structure without removal of H from PCN exhibits a much lower formation energy (∼0.15 eV) than Co–N4 and Co–N3 structures with removal of two H atoms from PCN (∼2.51 and 3.55 eV), demonstrating a high probability of existence of the Co–N4OH structure in BM-PCN/Co-c. This structure can also be evidenced by X-ray photoelectron spectroscopy (XPS). As shown in Fig. 2e, the Co 2p core-level XPS spectrum of BM-PCN/Co-c shows two distinct peaks at binding energies of 796.8 and 781.4 eV beside satellite peaks, corresponding to Co 2p1/2 and 2p3/2 of CoII ions.46 The spectrum of BM-PCN/Co also shows two Co 2p peaks but at a binding energy ∼1.1 eV higher, suggesting variation of the Co coordination structure from BM-PCN/Co to BM-PCN/Co-c. PCN/Co-c exhibits no peaks because of its low Co content. Fig. 2f shows O 1s core-level spectra of PCN, BM-PCN, BM-PCN/Co-c, and PCN/Co-c. All the samples exhibit one peak at a binding energy of ∼532.0 eV, ascribed to surface hydroxyl species,47 but an additional peak could be obtained for BM-PCN or BM-PCN/Co-c after deconvolution. The peak at a binding energy of ∼531.3 eV for BM-PCN should be assigned to adsorbed H2O at new active adsorption sites generated by ball milling. This peak can also be observed in the spectrum of BM-PCN/Co, but with a ∼0.1 eV shift to a higher binding energy (Fig. S11) owing to the influence of adsorbed CoII ions. The peak at ∼531.2 eV for BM-PCN/Co-c should be assigned to Co–OH,48 given that there is only one O 1s peak for BM-PCN-c (synthesized by direct calcination of BM-PCN) (Fig. S11). The calculated Co/O(–Co) molar ratio, based on the XPS data, is ∼1.07 (Table S2), close to 1, consistent with the Co–N4OH coordination structure.In C 1s and N 1s core-level XPS spectra, BM-PCN, BM-PCN/Co-c, PCN/Co-c, and BM-PCN/Co exhibit similar peaks to PCN (Fig. S12a–d), indicative of their similar framework structure which can also be evidenced by their similar N/C molar ratios, 1.53 (Table S3), but the N–H peak of BM-PCN shifts ∼0.2 eV to a lower binding energy relative to that of PCN, likely arising from the ball-milling induced destruction of intralayer hydrogen bonds (Fig. S13). The Co content in BM-PCN/Co, BM-PCN/Co-c, and PCN/Co-c is too low to cause detectable variation of C 1s and N 1s peaks. Similar FT-IR absorption bands of the samples (Fig. S14a and b) also indicate their basic frame structure, but in enlarged spectra (Fig. S14c), ν(C–N) and ν(C Created by potrace 1.16, written by Peter Selinger 2001-2019 N) absorption bands of BM-PCN shift 16 cm−1 to a higher wavenumber and 19 cm−1 to a lower wavenumber, respectively, relative to those of PCN at 1242 and 1640 cm−1,49 likely resulting from the ball-milling induced hydrogen bond destruction, and the shift of these two absorption bands turns smaller for BM-PCN/Co-c, suggesting calcination-induced reforming of the destroyed hydrogen bonds, which is consistent with the XPS results (Fig. S12c). Besides, BM-PCN exhibits a wider and relatively stronger ν(N–H)/ν(O–H) absorption band than PCN (Fig. S14a), probably owing to the hydrogen bond destruction and new adsorbed H2O, while this absorption band for BM-PCN/Co-c becomes much weaker, suggesting hydrogen bond reforming and loss of new adsorbed H2O (Fig. 2f). Zeta potentials of the samples dispersed in water reflect variation of surface adsorbed hydroxyl species. As shown in Fig. S15a, all the samples exhibit negative zeta potentials because of dissociation of surface hydroxyl species. The zeta potentials, following the order PCN (−24 mV) > BM-PCN (−41 mV) < BM-PCN/Co-c (−30 mV) ≈ PCN/Co-c (−28 mV), suggest the ball milling-induced increase of surface hydroxyls in BM-PCN and calcination-induced decrease in BM-PCN/Co-c, consistent with the FT-IR results.Solid-state 13C magic-angle-spinning nuclear magnetic resonance (NMR) spectra of PCN, BM-PCN, BM-PCN/Co-c, and PCN/Co-c show two similar peaks at chemical shifts of ∼164 and 156 ppm (Fig. S15b), ascribed to C−NHx and N Created by potrace 1.16, written by Peter Selinger 2001-2019 C–N, respectively,50 indicating their similar molecular framework, but in enlarged spectra, BM-PCN exhibits ∼0.3° movement of the N Created by potrace 1.16, written by Peter Selinger 2001-2019 C–N peak to a lower chemical shift compared with PCN, because of the ball-milling induced hydrogen bond destruction, and the C−NHx peak of BM-PCN/Co-c moves ∼0.2° to a lower chemical shift, likely owing to formation of the C–N–Co structure whose peak lies close to the C−NHx peak.51 The XRD patterns of the samples are shown in Fig. S15c. PCN and PCN/Co-c exhibit typical diffraction peaks of melon-type carbon nitride with a layered orthorhombic structure and peaks at 13.1° and 27.6° correspond to (210) and (002) facets, respectively,13,52 but BM-PCN reveals remarkably decreased peak intensity and ∼0.2° shift of the (002) peak to a lower 2θ (indicative of the increased interlayer distance) relative to PCN, demonstrating the ball-milling induced hydrogen bond destruction and substantial decrease of crystallinity. The remarkable decrease of crystallinity and almost no change of the surface area of BM-PCN, compared with those of PCN, further suggest that ball milling may form massive thin nanosheets (Fig. 1c) most of which stack into compact particles (Fig. 1b) owing to their high surface energy. In comparison with BM-PCN, BM-PCN/Co-c exhibits a narrower (002) peak, suggesting enhanced crystallinity owing to the calcination-induced hydrogen bond reforming, consistent with the FT-IR results. On the whole, it is likely the ball-milling induced destruction of hydrogen bonds that contributes largely to the increase of surface energy and new active adsorption centers and thus Co2+ adsorption on BM-PCN.Optical absorption capability of samples was investigated by UV-vis diffuse reflectance spectroscopy (DRS). As shown in Fig. 3a, BM-PCN/Co-c, BM-PCN, and PCN/Co-c exhibit considerably higher, lower, and similar optical absorption than/to PCN, respectively. For BM-PCN/Co-c, the optical absorption enhancement at a wavelength of <400 nm may benefit from the electron-rich Co that enhances π–π* transitions in heptazine rings,53 and the Urbach tail absorption should arise from the Co–OH doping.54,55 Bandgaps (Eg) of PCN, BM-PCN, BM-PCN/Co-c, and PCN/Co-c were roughly confirmed as 2.70, 2.81, 2.56, and 2.73 eV, respectively, via the formula Eg/eV = 1240/(λed/nm)56 where λed is the absorption edge determined by solid lines in the spectra. The wider Eg of BM-PCN probably results from the quantum size effect of massive ultrathin crystal nanosheets (Fig. 1c) formed by ball milling, and the narrower Eg of BM-PCN/Co-c arises from the Co–OH doping that was then verified by DFT calculations. As shown in Fig. S16, the calculated Eg of BM-PCN/Co-c, ∼1.90 eV, is much smaller than that of PCN (2.57 eV), in accordance with the experimental results. For PCN, the conduction band (CB) is contributed by C 2p and N 2p orbitals and the valence band (VB) mainly by N 2p orbitals, while for BM-PCN/Co-c, the CB is contributed by Co 3d, C 2p, and N 2p orbitals and the VB mainly by Co 3d and N 2p orbitals (Fig. S16c and d), effectively manifesting that the narrowing of Eg of BM-PCN/Co-c results from the Co–OH doping. In addition, there are prominent doping levels (Ed) in the bandgap of BM-PCN/Co-c, mainly contributed by Co 3d and O 2p orbitals (Fig. S16d), effectively proving the Co–OH doping effect in BM-PCN/Co-c. Similar calculation results have been reported for Pt–OH modified carbon nitride.57 Given that the experimental Co content (0.75 wt%) is much lower than the theoretical (6.71 wt%), practical doping levels in the bandgap may approach more to the VB. CB edges of the samples (ECB) could be roughly determined by using Mott-Schottky plots (Fig. S17) and their Fermi levels (Ef) were subsequently confirmed based on VB-XPS spectra (Fig. S18). Energy band levels of the samples are shown in Fig. 3b, and it seems that ball milling causes a slight downshift of the VB edge (EVB) of BM-PCN, favorable for photocatalytic water splitting, but the Co–OH doping causes a slight downshift of ECB and upshift of EVB of BM-PCN/Co-c. It is noteworthy that the Ed close to the VB edge (EVB) can capture photogenerated holes58 and thus the single-atom Co–OH works as the active site for the OER (Fig. 3b).Open in a separate windowFig. 3(a) UV-vis diffuse reflectance spectra of PCN, BM-PCN, BM-PCN/Co-c, and PCN/Co-c; (b) energy band levels of the samples and schematic illustration for water oxidation on BM-PCN/Co-c; (c) photoluminescence spectra, (d) time-resolved fluorescence spectra, and (e) anodic photocurrent (Ja) response of the samples; and (f) EPR spectra of the samples in the dark and under visible light irradiation. Data in (d) are the results of fitting decay curves to a tri-exponential model. Dark Ja in (e) was set as zero for distinct comparison.Spectroscopy and photoelectrochemical tests were conducted to evaluate photogenerated charge separation and transfer performance. As shown in Fig. 3c, photoluminescence (PL) spectra of all the samples show one emission peak, basically corresponding to their bandgap emission. BM-PCN exhibits weaker PL intensity than PCN, revealing a decreased photogenerated charge recombination efficiency, which originates from faster charge transfer from the inside to the surface of ultrathin nanosheets (Fig. S19) and trapped by surface states.59 BM-PCN/Co-c exhibits the lowest PL intensity and the PL intensity of PCN/Co-c is lower than that of PCN, which arises from the Ed capturing photogenerated holes to reduce their direct recombination with electrons beside the ultrathin nanosheet effect in BM-PCN/Co-c. Fig. 3d shows time-resolved fluorescence spectra of the samples. Decay curves were well fitted to a tri-exponential model (S3) and the obtained results are shown in Fig. 3d. Three lifetimes (τ1τ3) and their mean lifetime (τm, 89.2 ns) of BM-PCN are all much longer than those of PCN (τm = 17.9 ns), further suggesting the faster charge transfer from the inside to the surface of ultrathin nanosheets in BM-PCN, decreasing the direct charge recombination efficiency, but with subsequent surface radiative recombination.60 Interestingly, the τ1τ3 and τm (10.8 ns) of BM-PCN/Co-c are much shorter than those of PCN, which should result from faster transfer of holes to Ed that effectively decreases the charge recombination efficiency, with subsequent nonradiative energy transformation.61 The Co–OH doping effect also makes PCN/Co-c exhibit shorter τ1τ3 and τm (16.5 ns) than PCN. Fig. 3e shows the photocurrent response of the samples. Their anodic photocurrent density follows the order PCN < PCN/Co-c < BM-PCN < BM-PCN/Co-c, indicating gradually increased photogenerated charge separation efficiencies,62 basically consistent with the PL results. The relatively high photocurrent response of BM-PCN benefits from the applied bias that effectively inhibits surface recombination of photogenerated charge carriers.To assess charge mobility of the samples, their electrochemical impedance spectroscopy (EIS) spectra were tested with high-frequency data simply fitted to an equivalent circuit (Fig. S20). The obtained charge transfer resistance (Rct) follows the order PCN (26 Ω) > BM-PCN (18 Ω) ≈ PCN/Co-c (19 Ω) > BM-PCN/Co-c (13 Ω). Apparently, BM-PCN/Co-c exhibits smaller Rct than BM-PCN and PCN/Co-c, and PCN/Co-c exhibits smaller Rct than PCN, indicating the highest charge transfer performance of BM-PCN/Co-c63 which originates from the single-atom Co modification64 that may increase the electron density to facilitate charge transport. The smaller Rct of BM-PCN than that of PCN indicates the additional favorable effect of ultrathin nanosheets.65Fig. 3f shows electron paramagnetic resonance (EPR) spectra of the samples. All reveal one single Lorentzian line centered at a g of 2.0039, attributed to unpaired electrons in heptazine rings.66 In the dark, the EPR signal intensity follows the order PCN < BM-PCN < PCN/Co-c < BM-PCN/Co-c, and the stronger signal of BM-PCN than that of PCN results from formation of ultrathin nanosheets that enhances delocalization of unpaired electrons, while the stronger signal of BM-PCN/Co-c and PCN/Co-c mainly benefits from the Co doping that increases the delocalized electron density.67 Under visible light irradiation, the samples exhibit remarkable signal enhancement, following the sequence PCN < BM-PCN < PCN/Co-c < BM-PCN/Co-c, similar to that of the signal intensity in the dark, suggesting that the increase in the delocalized electron density facilitates charge photoexcitation. The high delocalized electron density favors charge transport, consistent with the EIS results, and the high photoexcited charge density benefits enhancement of photocatalytic activity.Photocatalytic OER activity of various samples was well evaluated using Ag+ as the sacrificial agent (Fig. S21). The Co content in BM-PCN/Co-c was optimized according to the photocatalytic OER rates and BM-PCN-c exhibits no detectable OER activity (Fig. S22), indicating indispensability of the Co–OH structure for the OER. The influence of the calcination temperature (Tc °C) of BM-PCN/Co on OER rates of BM-PCN/Co-c (Tc = 460) and BM-PCN/Co-cTc was investigated and BM-PCN/Co-c exhibits the highest photoactivity (Fig. 4a), manifesting that the optimal calcination temperature is 460 °C. Under both simulated solar light and visible light irradiation (λ ≥ 420 nm), BM-PCN/Co-c exhibits substantially higher OER activity than PCN/Co-c (Fig. 4b), further suggesting the significance of the single-atom Co loading amount, and remarkably higher activity than common PCN/CoOx (with 0.75 wt% Co, obtained via photodeposition) and BM-PCN-c/Co(OH)2 (with 0.75 wt% Co), demonstrating the high efficacy of the single-atom distribution of Co–OH in BM-PCN/Co-c. Besides, urea was used as the feedstock to synthesize carbon nitride (marked as PCN-urea) with a larger surface area (76 m2 g−1 (ref. 68)) than PCN, and PCN-urea was further used to synthesize PCN-urea/Co-c similar to the synthesis of BM-PCN/Co-c. The OER activity of BM-PCN/Co-c is prominently higher than that of PCN-urea/Co-c (with the optimized Co content and Co single atom distribution, Fig. S23), suggesting the significant role of ball milling in fabricating the single-atom Co–N4OH structure. To quantitively compare photoactivity of the samples, their mean OER rates under visible light illumination for 2 h are shown in Fig. 4c. The OER rate of BM-PCN/Co-c can reach ∼37.3 μmol h−1, about 13.8, 28.7, 2.6, and 2.0 times those of PCN/Co-c, PCN/CoOx, BM-PCN-c/Co(OH)2, and PCN-urea/Co-c, respectively. Comparatively, less N2 was generated for BM-PCN/Co-c (Fig. S24), further demonstrating the significance of single-atom Co–OH modification.Open in a separate windowFig. 4(a) The influence of the calcination temperature (Tc °C) of BM-PCN/Co on photocatalytic OER activity of BM-PCN/Co-c (Tc = 460) and BM-PCN/Co-cTc, under Xe-lamp illumination, with AgNo3 as the sacrificial agent; (b) photocatalytic oxygen evolution on various samples under Xe-lamp illumination with or without using a 420-nm filter; (c) corresponding OER rates of the samples in 2 h; (d) photocatalytic OER rates of BM-PCN/Co-c under irradiation with various monochromatic light sources for 12 h; (e) apparent quantum yields (AQYs) of BM-PCN/Co-c at different wavelengths and reaction times and the highest AQY at every wavelength, along with the UV-DRS spectrum; and (f) proposed mechanism for photocatalytic water oxidation on the single-atom CoII-OH structure.Photocatalytic oxygen evolution on BM-PCN/Co-c was also tested under monochromatic light irradiation (Fig. S25). Apparently, BM-PCN/Co-c can exhibit OER activity even at a wavelength of 500 nm. The mean OER rate in 12 h decreases from 1.85 to 0.54 μmol h−1 with increasing wavelengths from 400 to 500 nm (Fig. 4d), independent of light intensity of the Xe lamp and is mainly dependent on optical absorption capability of BM-PCN/Co-c at various wavelengths (Fig. 3a). Fig. 4e shows apparent quantum yields (AQYs) of BM-PCN/Co-c at different reaction times and wavelengths. Basically, there are maxima of AQYs with increasing reaction time at every wavelength, suggesting the adverse effect of excessive photodeposited Ag on surfaces of samples. These maxima are shown in Fig. 4e and accord well with the UV-vis DRS spectrum with increasing wavelengths. The maxima of AQYs at 400, 420, 450, and 500 nm can reach 4.69, 2.06, 1.07, and 0.46%, respectively. Compared with the reported photocatalytic OER results for PCN (Table S4), BM-PCN/Co-c exhibits the top-class performance.To investigate chemical stability of BM-PCN/Co-c, the cyclic OER experiment was conducted. After five consecutive runs, OER rates of BM-PCN/Co-c decrease less (Fig. S26a), with the morphology similar to the original (Fig. S26b). Co single atoms in the sample could still be distinctly observed by HAADF-STEM (Fig. S26c and d). In addition, N 1s core-level XPS spectra of BM-PCN/Co-c are almost similar before and after the cyclic experiment (Fig. S26e). These indicate the high stability of the basic framework structure of the sample. However, Co 2p core-level spectra show remarkable differences before and after the experiment, not only the CoII peak shift, probably owing to ion (e.g., IO4) adsorption, but also formation of a large amount of CoIII (Fig. S26f). Coexistence of CoII/CoIII may suggest the photocatalytic OER mechanism.The proposed OER mechanism based on the Co–OH structure is shown in Fig. 4f, according to the reported results in Mn doped PCN.34 Four holes are needed to complete four oxidation steps and obtain one O2 molecule. The first step starting with one hole may involve formation of the CoIII Created by potrace 1.16, written by Peter Selinger 2001-2019 O bond. The Co–N4OH structure should facilitate the water oxidation more compared with that of Co–N4 without OH coordination, by leaving out the initial adsorption process of H2O molecules.34 On the whole, the high photocatalytic OER activity of Co-PCN benefits from the Co–N4OH structure that not only effectively enhances optical absorption, and charge separation and transport, but also works as the highly active site for the OER.  相似文献   

13.
Mechanochemistry enables rapid access to boron imidazolate frameworks (BIFs), including ultralight materials based on Li and Cu(i) nodes, as well as new, previously unexplored systems based on Ag(i) nodes. Compared to solution methods, mechanochemistry is faster, provides materials with improved porosity, and replaces harsh reactants (e.g. n-butylithium) with simpler and safer oxides, carbonates or hydroxides. Periodic density-functional theory (DFT) calculations on polymorphic pairs of BIFs based on Li+, Cu+ and Ag+ nodes reveals that heavy-atom nodes increase the stability of the open SOD-framework relative to the non-porous dia-polymorph.

Mechanochemistry enables rapid access to boron imidazolate frameworks (BIFs), including ultralight materials based on Li and Cu(i) nodes, as well as new, previously unexplored systems based on Ag(i) nodes.

Mechanochemistry1–7 has emerged as a versatile methodology for the synthesis and discovery of advanced materials, including nanoparticle systems8–10 and metal–organic frameworks (MOFs),11–15 giving rise to materials that are challenging to obtain using conventional solution-based techniques.16–18 Mechanochemical techniques such as ball milling, twin screw extrusion19 and acoustic mixing20,21 have simplified and advanced the synthesis of a wide range of MOFs, permitting the use of simple starting materials such as metal oxides, hydroxides or carbonates,22,23 at room temperature and without bulk solvents, yielding products of comparable stability and, after activation, higher surface areas than solution-generated counterparts.24–29 The efficiency of mechanochemistry in MOF synthesis was recently highlighted by accessing zeolitic imidazolate frameworks (ZIFs)30,31 that were theoretically predicted, but not accessible under conventional solution-based conditions.17The advantages of mechanochemistry in MOF chemistry led us to address the possibility of synthesizing boron imidazolate frameworks (BIFs),32–34 an intriguing but poorly developed class of microporous materials analogous to ZIFs, comprising equimolar combinations of tetrahedrally coordinated boron(iii) and monovalent Li+ or Cu+ cations as nodes (Fig. 1A–C). Although BIFs offer an attractive opportunity to access microporous MOFs with lower molecular weights, particularly in the case of “ultralight” systems based on Li+ and B(iii) centers, this family of materials has remained largely unexplored – potentially due to the need for harsh synthetic conditions, including the use of n-butyllithium in a solvothermal environment.32–34Open in a separate windowFig. 1Structures of previously reported BIFs with: (A) zni-, (B) dia-, or (C) SOD-topology (M = Li, Cu); (D) tetrakis(imidazolyl)boric acids used herein for mechanochemical BIF synthesis; and (E) schematic representation of the herein developed mechanosynthesis of dia- and SOD BIF polymorphs based on Li, Cu or Ag metal nodes.We now show how switching to the mechanochemical environment enables lithium- and copper(i)-based BIFs to be prepared rapidly (i.e., within 60–90 minutes), without elevated temperatures or bulk solvents, and from readily accessible solid reactants, such as hydroxides and oxides (Fig. 1D and E). While the mechanochemically-prepared BIFs exhibit significantly higher surface areas than the solvothermally-prepared counterparts, mechanochemistry allows for expanding this class of materials towards previously not reported Ag+ nodes. The introduction of BIFs isostructural with those based on Li+ or Cu+ but comprising of Ag+ ions, enables a periodic density-functional theory (DFT) evaluation of their stability. This reveals that switching to heavier elements as tetrahedral nodes improves the stability of sodalite topology (SOD) open BIFs with respect to close-packed diamondoid (dia) topology polymorphs.As a first attempt at mechanochemically synthesis of BIFs, we targeted the synthesis of previously reported zni-topology LiB(Im)4 and CuB(Im)4 frameworks (Li-BIF-1 and Cu-BIF-1, respectively, Fig. 1A) using a salt exchange reaction between LiCl or CuCl with commercially available sodium tetrakis(imidazolyl)borate (Na[B(Im)4]) (Fig. 2A). Milling of LiCl and Na[B(Im)4] in a 1 : 1 stoichiometric ratio for up to 60 minutes led to the appearance of Bragg reflections consistent with the target Li-BIF-1 (CSD MOXJEP) and the anticipated NaCl byproduct. The reaction was, however, incomplete, as seen by X-ray reflections of Na[B(Im)4] starting material. In order to improve reactant conversion, we explored liquid-assisted grinding (LAG), i.e. milling in the presence of a small amount of a liquid phase (measured by the liquid-to-solid ratio η35 in the range of ca. 0–2 μL mg−1). Using LAG conditions with acetonitrile (MeCN, 120 μL, η = 0.5 μL mg−1) led to the complete disappearance of reactant X-ray reflections, concomitant with the formation of Li-BIF-1 alongside NaCl within 60 minutes.Open in a separate windowFig. 2(A) Reaction scheme for the mechanochemical synthesis of Li-BIF-1 by a salt metathesis strategy. Selected PXRD patterns for: (B) Na[B(Im)4] (C) LiCl, (D) simulated Li-BIF-1 (CSD MOXJPEP) and (E) synthesized BIF-1-Li by LAG for 60 minutes with MeCN (η = 0.5 μL mg−1), (F) CuCl, (G) simulated Cu-BIF-1 (CSD MOXJIT), and (H) synthesized BIF-1-Cu by LAG for 60 minutes with MeOH (η = 0.50 μL mg−1). Asterisks denote NaCl, a byproduct of the metathesis reaction. (Fig. 2B–E, also see ESI). The copper-based zni-CuB(Im)4 (Cu-BIF-1) was readily obtained from CuCl within 60 minutes using similar LAG conditions. We also explored LAG with methanol (MeOH), revealing that the exchange reaction to form NaCl took place with both LiCl and CuCl starting materials. With LiCl, however, the PXRD pattern of the product could not be matched to known phases involving Li+ and B(Im)4 (see ESI). With CuCl as a reactant, LAG with MeOH (η = 0.5 μL mg−1) cleanly produced Cu-BIF-1 alongside NaCl (see ESI).Next, we explored an alternative synthesis approach, analogous to that previously used to form ZIFs and other MOFs: an acid–base reaction between a metal oxide or hydroxide and the acid form of the linker: tetrakis(imidazolato)boric acid, HB(Im)4 (Fig. 3A).36–40 Neat milling LiOH with one equivalent of HB(Im)4 in a stainless steel milling assembly led to the partial formation of Li-BIF-1, as evidenced by PXRD analysis (see ESI). Complete conversion of reactants into Li-BIF-1 was achieved in 60 minutes by LAG with MeCN (η = 0.25 μL mg−1), as indicated by PXRD analysis (Fig. 3B–E), Fourier transform infrared attenuated total reflectance spectroscopy (FTIR-ATR), thermogravimetric analysis (TGA) in air, and analysis of metal content by inductively-coupled plasma mass spectrometry (ICP-MS) (see ESI).Open in a separate windowFig. 3(A) Reaction scheme for the mechanochemical synthesis of Li-BIF-1 using the acid–base strategy. Selected PXRD patterns for: (B) H[B(Im)4] (C) LiOH, (D) simulated Li-BIF-1 (CSD MOXJPEP), (E) synthesized BIF-1-Li by LAG for 60 minutes with MeCN (η = 0.25 μL mg−1), (F) Cu2O, (G) simulated Cu-BIF-1 (CSD MOXJIT), and (H) synthesized Cu-BIF-1 by ILAG for 60 minutes with MeOH (η = 0.50 μL mg−1) and NH4NO3 additive (5% by weight).Neat milling of HB(Im)4 with Cu2O under similar conditions gave a largely non-crystalline material, as evidenced by PXRD (see ESI). Switching to the ion- and liquid-assisted grinding (ILAG) methodology, in which the reactivity of a metal oxide is enhanced by a small amount of a weakly acidic ammonium salt, and which was introduced to prepare zinc and cadmium ZIFs from respective oxides,37–40 enabled the synthesis of Cu-BIF-1 from Cu2O. Specifically, PXRD analysis revealed complete disappearance of the oxide in samples obtained by ILAG with either MeOH or MeCN (η = 0.5 μL mg−1) in the presence of NH4NO3 additive (5% by weight, see ESI). Notably, achieving complete disappearance of Cu2O reactant signals also required switching from stainless steel to a zirconia-based milling assembly, presumably due to more efficient energy delivery.41 After washing with MeOH, the material was characterized by FTIR-ATR, TGA in air, and analysis of metal content by ICP-MS (see ESI).Whereas both the metathesis and acid–base approaches can be used to mechanochemically generate Li- and Cu-BIF-1, the latter approach has a clear advantage of circumventing the formation of the NaCl byproduct. Consequently, in order to further the development of mechanochemical routes to other BIFs, we focused on the acid–base strategy. As next targets, we turned to MOFs based on tetrakis(2-methylimidazole)boric acid H[B(Meim)4],36 previously reported32 to adopt either a non-porous diamondoid (dia) topology (BIF-2) or a microporous sodalite (SOD) topology (BIF-3) with either Li+ or Cu+ as nodes (Fig. 4). Attempts to selectively synthesize either Li-BIF-2 or Li-BIF-3 by neat milling or LAG (using MeOH or MeCN as liquid additives) with LiOH and a stoichiometric amount of HB(Meim)4 were not successful. Exploration of different milling times and η-values produced only mixtures of residual reactants with Li-BIF-2, Li-BIF-3, and/or not yet identified phases (see ESI). Consequently, we explored milling in the presence of 2-aminobutanol (amb), which is a ubiquitous component of solvent systems used in the solvothermal syntheses of BIFs.32,33 Gratifyingly, using a mixture of amb and MeCN in a 1 : 3 ratio by volume as the milling liquid led to an effective strategy for the selective synthesis of both the dia-topology Li-BIF-2 (CSD code MOXKUG), and the SOD-topology Li-BIF-3 (CSD code MUCLOM). The selective formation of phase-pure samples of Li-BIF-2 and Li-BIF-3 was confirmed by PXRD analysis, which revealed an excellent match to diffractograms simulated based on the previously reported structures (Fig. 4B–G). Systematic exploration of reaction conditions, including time (between 15 and 60 minutes) and η value (between 0.25 and 1 μL mg−1) revealed that the open framework Li-BIF-3 is readily obtained at η either 0.75 or 1 μL mg−1 after milling for 45 minutes or longer (Fig. 4B–G, also see ESI).§ Lower η-values of 0.25 and 0.5 μL mg−1 preferred the formation of the dia-topology Li-BIF-2, which was obtained as a phase-pure material upon 60 minutes milling at η = 0.5 μL mg−1, following the initial appearance of a yet unidentified intermediate. The preferred formation of Li-BIF-2 at lower η-values is consistent with our previous observations that lower amounts of liquid promote mechanochemical formation of denser MOF polymorphs.37Open in a separate windowFig. 4(A) Reaction scheme for the mechanochemical synthesis of Li-BIF-3. Comparison of selected PXRD patterns for the synthesis of Li-BIF-2 and Li-BIF-3: (B) H[B(Meim)4] reactant; (C) LiOH reactant; (D) simulated for Li-BIF-3 (CSD MUCLOM); (E) simulated for Li-BIF-2 (CSD MOXKUG); (F) Li-BIF-3 mechanochemically synthesized by LAG for 60 minutes with a 1 : 3 by volume mixture of amb and MeCN (η = 1 μL mg−1); and (G) Li-BIF-2 mechanochemically synthesized by LAG for 60 minutes with a 1 : 3 by volume mixture of amb and MeCN (η = 0.5 μL mg−1). Comparison of selected PXRD patterns for the synthesis of Cu-BIF-2 and Li-BIF-3: (H) Cu2O; (I) Cu-BIF-3 (CSD MOXJOZ); (J) Cu-BIF-2 (CSD MUCLIG); (K) Cu-BIF-3 mechanochemically synthesised by ILAG for 60 minutes using NH4NO3 ionic additive (5% by weight) and MeOH (η = 1 μL mg−1); and (L) mechanochemically synthesised Cu-BIF-2 by ILAG for 90 minutes using NH4NO3 ionic additive (5% by weight) and MeOH (η = 0.5 μL mg−1).Samples of both Li-BIF-2 and Li-BIF-3 after washing with MeCN were further characterized by FTIR-ATR, TGA in air, and analysis of metal content by ICP-MS (see ESI). Nitrogen sorption measurement on the mechanochemically obtained Li-BIF-3, after washing with MeCN and evacuation at 85 °C, revealed a highly microporous material with a Brunauer–Emmett–Teller (BET) surface area of 1010 m2 g−1 (Fig. 5A), which is close to the value expected from the crystal structure of the material (1200 m2 g−1, 32 For direct comparison with previous work,32 we also calculated the Langmuir surface area, revealing an almost 40% increase (1060 m2 g−1) compared to samples made solvothermally (762.5 m2 g−1) (Fig. 5A, inset).Experimental Brunauer–Emmett–Teller (BET) and Langmuir surface area (in m2 g−1) of mechanochemically synthesized SOD-topology BIFs, compared to previously measured and theoretically calculated values, along with average particle sizes (in nm) established by SEM and calculated energies (in eV) for all Li-, Cu-, and Ag-BIF polymorphs. The difference between calculated energies for SOD- and dia-polymorphs in each system is given as ΔE (in kJ mol−1)
MaterialSurface area (m2 g−1)Particle sizeb (nm)Electronic energy per formula unit (eV)ΔE (kJ mol−1)
Mechanochemical, BETMechanochemical, LangmuirPrior work, Langmuir 32Theoreticala
dia-Li-BIF-2−2679.17414.25
SOD-Li-BIF-310101060762.51200217 (n = 24)−2679.026
dia-Cu-BIF-2−3417.0919.67
SOD-Cu-BIF-39351196182.31100611 (n = 500)−3416.991
dia-Ag-BIF-2−4738.9598.66
SOD-Ag-BIF-3102012051170500 (n = 25)−4738.869
Open in a separate windowaCalculated using MOF Explorer (see ESI).bDetermined from SEM measurements, where n corresponds to number of particles observed.Open in a separate windowFig. 5BET adsorption plots for: (A) Li-BIF-3, showing a surface area of 1010 m2 g−1 and (B) Cu-BIF-3, showing a surface area of 935 m2 g−1. The insets in (A) and (B) are representative SEM images of the mechanochemically prepared BIF samples, with scale bars corresponding to 4 μm and 5 μm shown in white.The analogous copper(i)-based BIF-2 and BIF-3 frameworks were readily accessible by ILAG, by controlling the volume of the liquid additive and milling time (Fig. 4H–L, also see ESI). Similarly to our previous studies of ZIFs,17,24,37,39 increased milling times preferred the formation of the close-packed polymorph, dia-topology Cu-BIF-2. While the PXRD pattern of the reaction mixture after 60 minutes ILAG with MeOH (η = 0.5 μL mg−1) and NH4NO3 (5% wt/wt) indicated the presence of the SOD-topology Cu-BIF-3, longer milling led to the appearance of the dia-phase (see ESI). The materials were identified through comparison of experimental PXRD patterns to those simulated from published structures (CSD codes MUCLIG and MOXJOZ for Cu-BIF-2 and Cu-BIF-3, respectively).32 Quantitative synthesis of Cu-BIF-2 from Cu2O was readily accomplished by ILAG for 90 minutes (Fig. 4H–L). Following washing and drying, the products were characterized by PXRD, FTIR-ATR, TGA in air and ICP-MS elemental analysis of metal content.In order to achieve the synthesis of phase-pure microporous Cu-BIF-3, reaction conditions were modified by increasing η to 1 μL mg−1. This modification enabled the reproducible and quantitative synthesis of Cu-BIF-3 in 60 minutes milling (Fig. 4H–L), confirmed by PXRD, FTIR-ATR, TGA and elemental analysis of metal content (see ESI). Analyses by SEM and nitrogen sorption were performed on the mechanochemical product after washing and drying in vacuo at 85 °C, revealing that the sample consists of sub-micron particles and exhibits a high BET surface area of 935 m2 g−1, which is close to the theoretically expected value of 1100 m2 g−1 (Fig. 5B). To enable direct comparison with previously reported work,32 we also calculated the Langmuir surface area, revealing a 7-fold increase (1196 m2 g−1) compared to samples made solvothermally (182.3 m2 g−1) (), i.e. 34% lower compared to the mechanochemically synthesized sample, illustrating a clear benefit of mechanochemistry in providing a simpler, more efficient synthesis, as well as materials of improved porosity.32,42The mechanochemical approaches to Li- and Cu-based BIFs are surprisingly simple compared to previously reported solvothermal methods,32,33,42 not only avoiding bulk solvents and high temperatures (85 °C for Li-based, 120 °C for Cu-based BIFs), but also enabling the use of simple, easily handled solids LiOH and Cu2O as starting materials compared to, for example, n-BuLi.42 Notably, while the reported solvothermal synthesis of these materials also requires the use of amb for the preparation of both Li- and Cu-BIFs, the use mechanochemical conditions enabled amb-free synthesis of copper-based BIFs. Such simplifications of the synthetic procedure encouraged us to explore the possibility to extend this family of materials towards previously not reported silver(i) derivatives.As a starting material for the synthesis of Ag(i)-based BIFs we focused on Ag2CO3, generated in situ from readily accessible AgNO3 and K2CO3. One-pot milling reaction of HB(Meim)4, AgNO3, and K2CO3 in the respective stoichiometric ratios 1 : 1 : 1/2, using MeCN as the milling additive (η = 0.25 μL mg−1) readily produced the targeted AgB(Meim)4 material along with the side product KNO3 (Fig. 5, also see ESI). Specifically, analysis of the reaction mixtures by PXRD revealed that, similar to the lithium and copper(i) analogues,32 the silver-based BIF appears in two polymorphs which could be selectively synthesized by varying the milling time. The BIF products were readily separated from the KNO3 by-product after sequential washing with cold MeOH and acetone, and their respective structures were further validated by structure determination from PXRD data measured on washed and dried materials.Specifically, milling for 30 minutes led to the formation of a material (Ag-BIF-3) which, based on PXRD analysis, was isostructural to the SOD-topology Li-BIF-3 and Cu-BIF-3. Consequently, the crystal structure of Ag-BIF-3 (Fig. 5A) was determined through Rietveld refinement of a structural model based on the Cu-BIF-3 structure, in which the copper(i) sites have been replaced by Ag(i), giving rise to a cubic unit cell (space group P4̄3n as in the analogous Cu-BIF-3 and Li-BIF-3 structures) with a = 16.6659(3) Å. Composition of Ag-BIF-3 was verified by TGA/DSC and elemental analysis of metal content (see ESI). The microporous nature of the material was confirmed by N2 sorption analysis, which revealed a high BET surface area of 1020 m2 g−1. Sample analysis by SEM revealed dense aggregates of particles, with sizes below 100 nm (Fig. 5). The 13C cross-polarisation magic angle spinning (CP-MAS) solid-state nuclear magnetic resonance (ssNMR) spectrum of Ag-BIF-3 was consistent with the crystal structure, revealing three signals in the imidazolate region 100–160 ppm and the –CH3 group signal at ∼16 ppm (Fig. 6).Open in a separate windowFig. 6(A) Rietveld refinement of Ag-BIF-3 with difference plot shown in grey. (B) Rietveld refinement of Ag-BIF-2 with difference plot shown in grey. (C) BET adsorption plot Ag-BIF-3 showing a surface area of 1020 m2 g−1 and a SEM image of a representative sample (scale-bar 1 μm). Comparison of measured and simulated 13C CP-MA ssNMR spectra for silver-based BIFs: (D) calculated for Ag-BIF-3, (E) measured for Ag-BIF-3, (F) calculated for Ag-BIF-2 and (G) measured for Ag-BIF-2.Milling for 60 minutes under otherwise identical conditions led to a material whose PXRD pattern was very similar, but not identical, to that of dia-topology Li-BIF-2 and Cu-BIF-2 materials, with additional Bragg reflections indicating possible lower symmetry. The structure of this material (Fig. 6B) was determined by simulated annealing structure solution from PXRD data, revealing a monoclinic (space group P21) unit cell with a = 7.5198(4) Å, b = 16.3763(9) Å, c = 7.5876(4) Å and β = 90.136(6)o. In contrast to structures of Li-BIF-2 and Cu-BIF-2, which all exhibited one symmetrically independent Meim ligand in a tetragonal I4̄ space group, the structure of Ag-BIF-3 displays each tetrahedral node surrounded by four symmetrically non-equivalent imidazolate ligands. This much higher multiplicity is clearly reflected by the ssNMR spectrum of the material, validating the structure (Fig. 6). The composition of the material was similarly confirmed by TGA and by elemental analysis of the metal content (see ESI). For both Ag-BIF-2 and Ag-BIF-3 the measured 13C ssNMR chemical shifts were consistent with those calculated from the herein determined crystal structures (Fig. 6D–G). Notably, while materials based on silver(i) ions are often expected to be light sensitive, the herein reported Ag-BIF-2 and Ag-BIF-3 both appeared unchanged following six months exposure storage in a transparent vial on the bench.The crystal structures of Li-, Cu- and Ag-based BIFs provide a unique opportunity to evaluate the effect of changes in the metal node on the relative stability of BIF polymorphs with SOD- and dia-topology across three metals.43–45 The calculations were done using CASTEP plane-wave density-functional theory (DFT)46 code. The previously published crystal structures of Li- and Cu-BIFs with Meim linkers, as well as the structures of Ag-BIFs herein determined, were geometry-optimized using the PBE47 functional combined with many-body dispersion (MBD*)48–50 correction scheme. The PBE + MBD* approach has previously shown excellent agreement with experimental calorimetric measurements of ZIF polymorphs,24 therefore we expected the same approach to perform reliably for the structures of BIFs. In addition to calculating the relative energies of SOD- and dia-polymorphs, we have performed Gauge Including Projector Augmented Waves (GIPAW)51 simulation of the solid-state NMR spectra of Ag-BIFs to compare the simulated spectra with their experimental counterparts, confirming the low symmetry Ag-BIF-2 structure derived from PXRD data (Fig. 6D–G).Comparison of calculated energies reveals that increasing the atomic number of the metal node results in increased stabilization of the SOD-topology open framework with respect to the close-packed dia-polymorph. The energy differences (ΔE) between SOD- and dia-topology polymorphs for each pair of Li-, Cu-, and Ag-based frameworks are shown in 52–54The simulated ssNMR spectra of Ag-BIF-2 and Ag-BIF-3 showed excellent agreement with the experiment (Fig. 6) in terms of overall chemical shift and the number of distinct NMR signals arising from the crystallographic symmetry. The spectrum of the SOD polymorph is consistent with a single symmetrically unique Meim linker, while the signal splitting found in the spectrum of the dia-polymorph corresponds to four distinct 2-methylimidazolate units. The NMR simulation fully supports the structural models derived from PXRD data, with calculated chemical shifts underlining the accuracy of the herein used theoretical approach.  相似文献   

14.
HydroFlippers are introduced as the first fluorescent membrane tension probes that report simultaneously on membrane compression and hydration. The probe design is centered around a sensing cycle that couples the mechanical planarization of twisted push–pull fluorophores with the dynamic covalent hydration of their exocyclic acceptor. In FLIM images of living cells, tension-induced deplanarization is reported as a decrease in fluorescence lifetime of the dehydrated mechanophore. Membrane hydration is reported as the ratio of the photon counts associated to the hydrated and dehydrated mechanophores in reconvoluted lifetime frequency histograms. Trends for tension-induced decompression and hydration of cellular membranes of interest (MOIs) covering plasma membrane, lysosomes, mitochondria, ER, and Golgi are found not to be the same. Tension-induced changes in mechanical compression are rather independent of the nature of the MOI, while the responsiveness to changes in hydration are highly dependent on the intrinsic order of the MOI. These results confirm the mechanical planarization of push–pull probes in the ground state as most robust mechanism to routinely image membrane tension in living cells, while the availability of simultaneous information on membrane hydration will open new perspectives in mechanobiology.

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

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

15.
We present a new operando approach for following reactions taking place in mechanochemistry, relying on the analysis of the evolution of the sound during milling. We show that differences in sound can be directly correlated to (physico)chemical changes in the reactor, making this technique highly attractive and complementary to others for monitoring mechanochemical reactions. Most notably, it can provide unique information on the actual movements of the beads within the milling jars, which opens new avenues for helping rationalize mechanochemical processes.

We present a new operando approach for following reactions taking place in mechanochemistry, relying on the analysis of the evolution of the sound during milling.

The growing attention to mechanochemistry has led to an impressive range of research areas to which this “green” synthetic method can be applied, ranging from pharmacy to catalysis as well as crystal engineering or materials science.1,2 Although widely used for synthesizing elaborate compounds, the details of grinding and more specifically of the physical and chemical mechanisms taking place are still often seen as unfathomable. In the case of ball-milling, this is mainly due to the opaque and closed environment of the milling reactors in which most reactions take place.In light of the great interest in mechanochemistry, several research groups have thus strongly focused their attention on developing in situ as well as operando analytical methods in order to monitor reactions and observe evolutions in the structure, crystallinity and/or texture of compounds involved during the milling processes.3–8 Except for temperature, these methods usually require the use of transparent reactor materials (for Raman or X-ray diffraction monitoring, for example),4,9 or the adaptation of the geometry and size of the reactor itself (in particular in the case of NMR analyses).10–16 However, several studies have shown that the kinetics and/or reaction mechanisms in ball-milling can be greatly affected by the size and materials composing the reactor and beads, meaning that the latter methods cannot be readily applied to different reactions without re-investigating the synthetic conditions.17,18 Hence, developing new operando approaches allowing the evolution of reaction media to be followed, but without requiring any modification of the initial set of experimental parameters, would be greatly beneficial for the mechanochemistry community.Several mechanical behaviours involving the beads and reactors have been shown to come into play during a ball-mill synthesis.19–22 The reacting particles undergo direct impacts from the beads, as well as shearing forces, which both can lead to particle size reduction and the creation of “fresh reactive surfaces”, and thereby enable the chemical reactions to take place. Moreover, under some milling conditions, specific trajectories of the beads have also been identified, such as their direct back and forth movements from one wall of the reactor to another, or their rolling along the sides of the milling reactor. Such processes are generally difficult to predict, and depend on numerous experimental parameters such as the type of beads (size, number and material), the type of reactors (volume, geometry, and material), the milling parameters (frequency, time), not to mention the type of movements imposed to the reactors by the mill (e.g., planetary, oscillating vertically or horizontally), but also the physical and chemical properties of the milled reagents. Various sounds can sometimes be noticed during the milling of compounds, which attest of changes in the beads'' motions in the reactor. Yet, although a few previous attempts to follow the sound during ball-milling experiments had been reported, notably in 2004 by Cuadrado-Laborde et al.,23 to the best of our knowledge, only sounds resulting from the impact of a bead on the reactor were identified and analysed. No systematic investigation was made at the time on how the sound recorded could inform on other movements of the bead (aside from the simple impact), nor, more importantly, on how it could be used to directly follow different physico-chemical changes occurring in the milling reactors.23 In this context, we decided to record the sound produced by a vertical mixer mill upon milling, in order to try to decipher the acoustic variations arising from different movements of the beads, and how they can be related to changes in the composition or physico–chemical properties of the medium. Moreover, as detailed below, we investigated how this method compares to other known operando techniques, i.e. Raman spectroscopy and temperature measurements, for following reaction mechanisms taking place during ball-milling (see Fig. 1).Open in a separate windowFig. 1Top: schematic representation of the set-up used for the operando monitoring of mechanochemical reactions, using the combination of Raman spectroscopy, thermal imaging, and the recording of an AF spectrogram. Bottom: photo of the instrumental set-up used in this work.As a first system of study, we decided to follow the mechanochemical formation of a 1 : 1 cocrystal of terephthalic acid (H2BDC) and 1,4-diazabicyclo[2.2.2]octane (DABCO) (see Fig. 2a). The sound was recorded through the use of a microphone situated close to the milling reactor. This recording was then used to generate an audio-frequency spectrogram (AF spectrogram), which allows the Fourier transform of the signal to be monitored as a function of time, and different sound harmonics to be observed. In parallel to the sound recording, Raman analyses and temperature measurements were also carried out to follow the evolution of the reaction medium (Fig. 1). These three operando methods were adapted to a vertical mixer mill (Pulverisette 23, Fritsch) using a Perspex reactor with two ZrO2 beads (Ø 10 mm). Here, the experiments were conducted at a frequency of 50 Hz with a reaction time of 60 min, for a total mass of reactants ∼500 mg (see ESI for more details).Open in a separate windowFig. 2(a) Dry milling reaction between H2BDC (1.0 eq) and DABCO (1.0 eq) during 60 min at 50 Hz, inducing the formation of a cocrystal (see Fig. S1 and S2 for ex situ pXRD and Raman analyses of the cocrystal). (b). Left: intensity of the 3rd harmonic of the AF spectrogram as a function of time, obtained by integration between 145 and 155 Hz. Centre: temperature profile of the reactor during the milling reaction presented in (a). Right: time-resolved operando Raman spectra for a selected area (between 950 and 1050 cm−1); the intensity scale goes from low (blue) to high (red and then white). The red shaded rectangle emphasizes the presence of the intermediate species, while purple ones highlight the erratic rolling motions.During the mechanochemical formation of the H2BDC–DABCO cocrystal (Fig. 2a),24 various sounds were noticed. Careful analysis of the AF spectrogram revealed contrasts in signal intensity in the 105–145 Hz and 145–155 Hz regions (see Fig. S3b), i.e. close to the 3rd harmonic of the milling frequency (50 Hz in this case). Notably, the acoustic intensity between 145 and 155 Hz was found to significantly increase between ∼1000 and 1500 s of milling (Fig. 2b, red shaded zone). This experimental observation could be reproduced in the same time-frame by repeating the experiment (see Fig. S4). Interestingly, when looking at the measurements recorded using the other operando methods, an increase of the reactor''s temperature by about 10–12 °C was simultaneously observed using the thermal imaging camera (Fig. 2b). Moreover, the Raman spectra recorded over the same period of time also revealed shifts in the DABCO and H2BDC vibration bands (Fig. 2b). Hence, these first results suggest that changes in sound can be directly correlated to chemical reactions occurring in the reactor.Further analysis of the Raman spectra was performed (Fig. S5), focusing notably on the zone between 950 and 1050 cm−1, in which the DABCO νCC/ωCH2 vibration bands appear.25 The latter vibration mode was first found to shift from 970.8 cm−1 to 984.3 cm−1 after ∼1000 s of milling, and then to further shift to 1013.5 cm−1 (after ca. 500 s), which corresponds to the vibration frequency expected for the final cocrystal (Fig. 2b). Although the exact nature of the intermediate has not yet been unambiguously identified, it can be tentatively assigned to an {H2BDC–DABCO} species, in which only one of the nitrogen atoms of DABCO becomes protonated (based on the shift observed in Raman spectroscopy),25 leading to a single N–H⋯O hydrogen-bonding motif (in contrast with the final cocrystal which contains two). Remarkably, the previously noticed variation of sound around 150 Hz (3rd harmonic) happens precisely during the short lifespan of this intermediate, i.e. between ∼1000 and 1500 s. To the best of our knowledge, it is the first time that evidence of the presence of reaction intermediates in mechanochemistry is accounted for by analysing the sound occurring during the milling.In order to help rationalize the origin of the changes in sound during the formation of the intermediate, the movements of the beads across the transparent Perspex milling reactors were visualized using a digital camera. In doing so, it was found that when the intermediate appears (thus when the intensity of the sound is maximum in the 145–155 Hz frequency range, i.e. in the red-shaded area on Fig. 2b), the beads undergo a rolling type of movement on the wall of the reactor. Such motion can be opposed to the erratic impact of the beads which occurs during most of the remaining time, and can be identified here on the AF spectrogram by the more intense inharmonic signals between 105 and 145 Hz (see ESI Fig. S3). It is possible that the predominance of rolling between 1000 and 1500 s may be caused by the significant changes in texture of the medium upon formation of the intermediate. More generally speaking, to the best of our knowledge, it is the first time that an acoustic signature characteristic of rolling events in mechanochemistry is identified through sound measurements.Beyond the possibility of directly observing the formation and/or disappearance of a reaction intermediate, we also found that the analysis of the AF spectrogram actually helps understand deeper some of the details of the data recorded using Raman spectroscopy and the thermal imaging camera. For this reaction system, the “rolling” vs. “impact” sound signatures were found to largely appear in specific regions, as illustrated in Fig. S6. Indeed, the impact is appearing in the 105–145 Hz region, while rolling is visible in the 145–155 Hz one. Yet, interestingly, the signature characteristic of rolling motions was observed not only during the life-span of the intermediate, but also intermittently before and after its appearance (see for example the purple-shaded rectangles in Fig. 2b), with a simultaneous decrease in the Raman signals intensities and increase in temperature. Concerning the Raman data, this can be explained by the fact that when rolling occurs, the beads are more likely to remove the powder from the point of focus of the Raman laser, implying that less compound is then analysed during the Raman acquisition time, hence the decrease in signal intensity. Regarding temperature, the small increase may be caused by a more significant frictional heating of the beads against the reactor''s walls during the rolling events.26Overall, throughout this first example, we have been able to show how sound modifications can attest of a chemical evolution of the medium inside the milling reactor in a complementary way to two other operando techniques (Raman analyses and thermal imaging measurements), and how they can also explain some of the smaller features observed during the acquisition of the operando data (notably when rolling motions occur). In light of these first results, we decided to study on one hand whether other changes in the physical aspect of a reaction medium could be correlated to variations in sound, and on the other hand how sound measurements may actually inform on chemical events taking place in the milling reactors, in situations when temperature measurements or Raman operando analyses can be ambiguous, poorly informative, or not even applicable.In order to investigate how changes in physical aspect of a reaction medium can be followed through the recording of AF spectrograms, we decided to study the hydrolysis of SiO2 (initially introduced in the reactor as “fumed silica”), upon progressive addition of small amounts of water during ball-milling. Indeed, as shown in Fig. 3, this system appeared to us as ideal for such investigations, because it goes through many of the physical states commonly encountered in ball-milling reactions. First, the medium is under the form of a light fluffy powder, which upon addition of water forms more agglomerated particles (leading to a “heavy powder”), which then further agglomerate and stick to the beads (leading to a so-called “snow-ball” effect).7,8,27,28 The subsequent addition of water enables the medium to re-spread onto the walls of the reactor (the phase then having the aspect of a “sticky gel” or “gel”), and finally evolves towards a slurry type of texture. These changes were thus followed using a similar set-up as for the study of the H2BDC–DABCO cocrystal (P23 vertical mixer mill operated at 50 Hz, using a Perspex reactor with two zirconia beads), and after each addition of water, the milling was performed for 5 minutes. Here, in addition to recording the AF spectrogram, the changes in temperature of the reactor were monitored using a thermal-imaging camera, and the movements of the beads inside the reactor were also filmed at specific time-points using a digital camera. It is worth noting that the observations described below were found to be essentially reproducible (see Fig. S7 for comparison of AF spectrograms).Open in a separate windowFig. 3(a) Scheme of the hydrolysis of SiO2 in a Perspex reactor using two zirconia beads. After each addition of 10 μL of H2O, the mixture was milled during 5 min at 50 Hz. (b) Sound intensity variation of the 3rd harmonic (145–155 Hz) according to the physical state of SiO2 upon addition of H2O. Associated pictures are taken just before the water addition. The red dashed line is given as a guideline to show the differences in sound intensities for motions which are predominantly impact. The blue arrows indicate the 5 minutes periods of milling when impact motions were predominant. (c) Temperature profile of the reactor. Blue shaded rectangles emphasize the impact motions regions. The green arrows highlight the regions which show similar temperature profiles (but different sound signatures). (d) Details of the three different bead motions detected during the experiment (on the left: pictures of the reactor, on the right: schemes of the beads motion): impact, rolling under snow-ball effect and rolling.When looking at the full AF spectrogram, it is clear that it is around the third harmonic (∼150 Hz) that the contrast is the most significant and the intensity the highest (see Fig. S8 in ESI for changes in sound intensity in other spectral regions). As for the H2BDC–DABCO system, it was found that high intensity signals around this frequency can attest of rolling-types of motions, as shown for example by analysing the results in the zone corresponding to the addition of 180–300 μL of water, in which the presence of rolling is clearly visible when visualising the movement of the beads and looking at the aspect of the medium after opening of the reactor. Concomitantly, the events with the lowest intensity signals at ∼150 Hz (i.e. after addition of 140 and 170 μL) could be corroborated with more “impact-like” movements (see blue arrows in Fig. 3b, zone corresponding to the “sticky gel” aspect of the mixture). These events were also found to correspond to a simultaneous increase in signal intensity in the 105–145 Hz region, in line with previous observations made for the H2BDC–DABCO system (see Fig. S9). However, it is important to highlight that in the case of the hydrolysis of silica, a lower intensity signal at ∼150 Hz did not systematically correlate with a higher intensity sound in the 105–145 Hz region (see ESI, Fig. S8), in contrast with what had been seen previously for the cocrystal (see below for further discussion). This shows that in order to help interpret the subtleties of the movements of the beads within the reactor and of the physico–chemical processes taking place, the investigation of several zones of the AF spectrograms is actually needed.When analysing in more detail the overall plot of the changes in sound intensity around the 3rd harmonic as a function of the amount of water added (Fig. 3), it is clear that the most significant increase occurs in the zone corresponding to the formation of the “snow-ball” (i.e. after addition of 90–130 μL of water – red-shaded zone), when the beads start rolling against the reactor''s walls and accumulating powder at their surface. Most interestingly, when looking at the movement of the beads in this regime using a digital camera, a rolling of both beads in separate zones of the reactor could be observed, as tentatively illustrated in Fig. 3d. A possible explanation to the related increase in sound would be that upon the formation of the snow-balls, the walls of the reactor then become mostly free of powder, inducing such sound intensification. A contrario, when the powder is homogeneously spread over the walls of the reactor, as is the case upon formation of the “heavy powder” (zones highlighted in light blue in Fig. 3, which correspond essentially to rolling motions), the sound intensity is much more dampened (see Fig. 3, S7 and S9).The importance of being able to follow the sound during the milling further appears when comparing the results of the AF spectrogram analyses to those of the temperature measurements. In several situations, just like for the H2BDC–DABCO system, the rolling events were found to be associated to temperature increases inside the milling reactor, as shown in particular after addition of 150 μL and of 180 to 250 μL of water. However, changes in sound intensity around 150 Hz did not systematically correlate in the same way to changes in temperature. For example, while the temperature profiles after addition of 70 and 110 μL of water were essentially the same (see green arrows on Fig. 3c), the sound intensity at ∼150 Hz over that same period of time was very different, in one case corresponding to the milling of a “heavy powder”, while in the other to a snowball. Such results clearly demonstrate the complementarity of acoustic and temperature measurements in trying to decipher the evolution in the physico–chemical properties of the medium during the milling. Considering the increasing number of studies in which temperature measurements are reported,26,29,30 and related to endo- or exothermal chemical reactions occurring in the milling reactors, being able to take into consideration that changes in beads motions can also be correlated to changes in temperature is another important feature to bear in mind.Overall, although many other aspects related to the milling of silica in presence of water would deserve to be investigated, these results on the hydrolysis of silica clearly highlight that (i) beyond the identification of rolling or erratic impact movements of the beads, acoustic measurements can attest of other phenomena occurring in the reactor, such as the formation of snow-balls (during which a different type of rolling movements take place), or the evolution of the aspect of the mixture (e.g. formation of a “sticky gel” inducing impact motions); and (ii) sound measurements are highly complementary to temperature measurements, as they can bring unique evidence of phenomena taking place during the milling.Finally, in order to demonstrate the utility of sound measurements for the optimization of ball-milling reaction conditions, we decided to study the hydrolysis of an activated fatty acid, lauric acid, using this operando method. Indeed, during our previous work on the 17O/18O isotopic enrichment of fatty acids,31,32 we had found that for saturated fatty acids, the hydrolysis time of the activated acyl-imidazole intermediates had to be systematically optimized by repeating the experiments with various milling times. Given that this hydrolysis is accompanied by a change in texture of the reaction medium, which switches from a powdery aspect to a more oily one, we suspected that acoustic measurements could be used to determine the time after which the hydrolysis is complete (Fig. 4a).Open in a separate windowFig. 4(a) Reaction corresponding to the mechanochemical hydrolysis of activated lauric acid (CH3–(CH2)10–COOH), which was followed here using operando analyses. (b) Operando experiments conducted in a Raman laser-transparent Perspex reactor (see picture in the orange rectangle, left) and in an opaque stainless steel reactor (see picture in the blue rectangle, right) with: temperature profiles of the reactor during the milling reactions, intensity of the 5th harmonic of the AF spectrograms as a function of time obtained by an integration between 205 and 245 Hz, and Raman analyses (for the experiment in a Perspex reactor). Dotted red lines are used as visual helps for indicating the completion of the reaction.Two experimental conditions were compared: the first hydrolysis experiment was performed in a Perspex reactor with two zirconia beads (10 mm Ø) in order to allow the Raman signals to be recorded, while the second one was performed in a stainless steel reactor with two stainless steel beads (10 mm Ø), and can be seen as a “blind” experiment (as it cannot be followed by Raman spectroscopy). The changes in sound, temperature, and, in the former case, in Raman signatures, during the milling are compared in Fig. 4.For these milling experiments, carried out at a frequency of 50 Hz, it was found that the highest contrast on the AF spectrograms appears close to the 5th harmonic, and more specifically in the 205–245 Hz frequency region (see ESI Fig. S10–S12). For both types of milling reactors, after an initial period during which significant variations in the sound intensity are observed, there is a stabilisation of the sound intensity (see dashed red vertical line in Fig. 4b), suggesting that a final physical state of the reaction medium has been reached. When looking at the simultaneous temperature measurements, it appears that this time also corresponds to the threshold after which no strong changes in reactor''s temperature are noticed. Moreover, when looking at the operando Raman data recorded for the experiment carried out in a Perspex reactor, it is clear that this threshold also corresponds to the period after which no further change in the vibration frequencies occurs, as illustrated by focusing on the 1100–1200 cm−1 spectral region. Using ex situ analyses by FTIR spectroscopy, it was verified that after stabilization of the sound the hydrolysis is indeed complete, for reactions carried out both in Perspex and stainless steel reactors (see Fig. S12). Overall, this demonstrates the utility of performing sound measurements in operando mode, in order to determine the time at which experiments are over, especially in situations when simultaneous Raman analyses are not possible, for example when working in a stainless steel reactor, or when temperature measurements may not be possible, for example in cases when dedicated temperature-regulated milling systems are used for the experiments.In conclusion, we have presented here a new approach based on the study of the sound for following mechanochemical syntheses using operando methods. The potential of this approach was illustrated in three different examples, demonstrating how it can be used to draw attention to the formation of synthetic intermediates, to know when a reaction is complete, and also to follow changes in the physical state of a reaction medium. Moreover, it was shown that sound analyses can shed light on different types of movements of the beads within the milling reactors, meaning that it is highly interesting to use such analyses to decipher the different phenomena taking place within the milling jars. Overall, this makes sound measurements a promising addition to the toolbox of methods which mechanochemists can now use to try to understand and control the evolution of reaction media during ball-milling.  相似文献   

16.
Pyrrolidine, an important feedstock in the chemical industry, is commonly produced via vapor-phase catalytic ammoniation of tetrahydrofuran (THF). Obtaining pyrrolidine with high purity and low energy cost has extremely high economic and environmental values. Here we offer a rapid and energy-saving method for adsorptive separation of pyrrolidine and THF by using nonporous adaptive crystals of per-ethyl pillar[6]arene (EtP6). EtP6 crystals show a superior preference towards pyrrolidine in 50 : 50 (v/v) pyrrolidine/THF mixture vapor, resulting in rapid separation. The purity of pyrrolidine reaches 95% in 15 min of separation, and after 2 h, the purity is found to be 99.9%. Single-crystal structures demonstrate that the selectivity is based on the stability difference of host–guest structures after uptake of THF or pyrrolidine and non-covalent interactions in the crystals. Besides, EtP6 crystals can be recycled efficiently after the separation process owing to reversible transformations between the guest-free and guest-loaded EtP6.

Here we offer a rapid and energy-saving method for adsorptive separation of pyrrolidine and tetrahydrofuran by using nonporous adaptive crystals of per-ethyl pillar[6]arene.

Pyrrolidine is an important feedstock in the chemical industry that has been widely used in the production of food, pesticides, daily chemicals, coatings, textiles, and other materials.1 Particularly, pyrrolidine is a raw material for organic synthesis of medicines such as buflomedil, pyrrocaine, and prolintane.2 Moreover, pyrrolidine is also used as a solvent in the semi-synthetic process of simvastatin, one of the best-selling cardiovascular drugs.3 In the chemical industry, there are many preparation methods for pyrrolidine. The most common way to obtain pyrrolidine is the gas-phase catalytic method using tetrahydrofuran (THF) and ammonia as raw materials;4 this is carried out at high temperature under catalysis by solid acids. However, separating pyrrolidine from the crude product is difficult because of similar molecular weights and structures between pyrrolidine (b.p. 360 K and saturated vapor pressure = 1.8 kPa at 298 K) and THF (b.p. 339 K and saturated vapor pressure = 19.3 kPa at 298 K), which result in complicated processes and large energy consumption.5 Therefore, it is worthwhile to find energy-efficient and simple methods to separate pyrrolidine from THF.Many techniques and materials, including porous zeolites, metal–organic frameworks (MOFs), and porous polymers, have facilitated energy-efficient separations of important petrochemicals and feedstocks, including THF and pyrrolidine.6,7 However, some drawbacks of these materials cannot be ignored.8 For example, the relatively low thermal and moisture stabilities of MOFs limit their practical applications. Therefore, the development of new materials with satisfactory chemical and thermal stabilities for pyrrolidine/THF separation is of high significance.In the past decade, pillararenes have been widely studied in supramolecular chemistry.9 Owing to their unique pillar structures and diverse host–guest recognitions, pillararenes have been used in the construction of numerous supramolecular systems.10 Recently, nonporous adaptive crystals (NACs) of macrocycles, which have shown extraordinary performance in adsorption and separation, have been developed by our group as a new type of adsorption and separation materials.11 Unlike MOFs, covalent-organic frameworks (COFs), and other materials with pre-existing pores, NACs do not have “pores“ in the guest-free form, whereas they adsorb guest vapors through cavities of macrocycles and spaces between macrocycles. NACs have been applied in separations of many significant chemicals such as alkane isomers, aromatics, and halohydrocarbon isomers.12 However, such materials have never been used to separate pyrrolidine and THF. Herein, we utilized pillararene crystals as a separation material and realized the selective separation of pyrrolidine from a mixture of pyrrolidine and THF. We found that nonporous crystals of per-ethyl pillar[6]arene (EtP6) exhibited a shape-sorting ability at the molecular level towards pyrrolidine with an excellent preference, while crystals of per-ethyl pillar[5]arene (EtP5) did not (Scheme 1). In-depth investigations revealed that the separation was driven by the host–guest complexation between pyrrolidine and EtP6, which resulted in the formation of a more stable structure upon adsorption of pyrrolidine vapor in the crystalline state. EtP6 crystals can also adsorb THF. However, when these two chemicals simultaneously exist as the vapor of a 50 : 50 (v/v) mixture, EtP6 prefers pyrrolidine as an adsorption target. Compared with previously reported NAC-based separation, this separation took place rapidly. 95% purity was achieved in 15 min, and the purity increased to 99.9% after 2 h of separation. Moreover, pyrrolidine was removed upon heating, along with the structural transformation of EtP6 back to its original state, endowing EtP6 with excellent recyclability.Open in a separate windowScheme 1Chemical structures and cartoon representations: (a) EtP5 and EtP6; (b) THF and pyrrolidine.EtP5 and EtP6 were prepared as previously described and then a pretreatment process was carried out to obtain guest-free EtP5 and EtP6 (Fig. S1–S4†).13 According to powder X-ray diffraction (PXRD) patterns, activated EtP5 and EtP6 (denoted as EtP5α and EtP6β, respectively) were crystalline, and the patterns matched previous reports (Fig. S5 and S6).14 Studies from our group indicated that EtP5α and EtP6β crystals were nonporous, presumably due to their dense packing modes.We first investigated the adsorption capabilities of EtP5α and EtP6β towards pyrrolidine and THF vapors. Based on time-dependent solid–vapor adsorption procedures, both EtP5α and EtP6β showed good ability to adsorb pyrrolidine and THF vapors. As shown in Fig. 1a, the adsorption amount of THF in EtP5α was higher than that of pyrrolidine. It took 6 hours for EtP5α to reach saturation points for adsorption of both pyrrolidine and THF vapors. The final storage of THF in EtP5α was 2 : 1 (molar ratio to the host), whereas the storage of pyrrolidine was 1 : 1. It seemed that the THF vapor was favored to occupy EtP5α, which was ascribed to the relatively lower boiling point of THF. A similar phenomenon was found for EtP6β. Time-dependent solid–vapor adsorption experiments for pyrrolidine demonstrated that it took just 1 hour to reach the saturation point, while it took 4 hours for the THF vapor (Fig. 1b). The adsorption amount of THF vapor was twice that of pyrrolidine. 1H NMR spectra and thermogravimetric analyses (TGA) further confirmed the adsorption and storage of THF and pyrrolidine in both hosts (Fig. S7–S16†). Meanwhile, in the desorption process, adsorbed pyrrolidine and THF in EtP6β were easily released under reduced pressure and heating. Based on these data, it was clear that pyrrolidine could be adsorbed rapidly by both EtP5α and EtP6β in molar ratios = 1 : 1, while THF could be captured in a relatively slow process. Structural changes after adsorption of these two vapors were analyzed via PXRD experiments, in which varying degrees of changes before and after adsorption were observed, evidencing the appearance of new crystal structures (Fig. 1c and d). Nevertheless, only slight differences were observed in the PXRD patterns after the adsorption of THF or pyrrolidine, which might be ascribed to the structural similarity of the two molecules.Open in a separate windowFig. 1Time-dependent solid–vapor adsorption plots of (a) EtP5α and (b) EtP6β for single-component pyrrolidine and THF vapors. PXRD patterns of (c) EtP5α and (d) EtP6β: (I) original activated crystals; (II) after adsorption of THF vapor; (III) after adsorption of pyrrolidine vapor.To study the mechanism of adsorption, guest-loaded single crystals were obtained by slowly evaporating either THF or pyrrolidine solutions of pillararenes (Tables S2 and S3). In the crystal structure of THF-loaded EtP5 (2THF@EtP5, Fig. 2a and S17),11a two THF molecules are in the cavity of one EtP5 molecule driven by multiple C–H⋯O hydrogen bonds and C–H⋯π bonds. EtP5 assembles into honeycomb-like infinite edge-to-edge 1D channels. In the crystal structure of pyrrolidine-loaded EtP5 (pyrrolidine@EtP5, Fig. 2b and S19), one pyrrolidine molecule, stabilized by C–H⋯π interactions and C–H⋯O hydrogen bonds between hydrogen atoms on pyrrolidine and oxygen atoms on EtP5, is found in the cavity of EtP5. It''s worth mentioning that a hydrogen atom which is linked with the N atom of pyrrolidine also forms a strong hydrogen bond with an oxygen atom on the ethoxy group of EtP5. EtP5 forms imperfect 1D channels because of partial distortion of orientation. The PXRD patterns simulated from these crystal structures matched well with the experimental results (Fig. S18 and S20), which verified that the uptake of vapors transformed EtP5α into pyrrolidine-loaded EtP5.Open in a separate windowFig. 2Single crystal structures: (a) 2THF@EtP5; (b) pyrrolidine@EtP5.In the crystal structure of THF-loaded EtP6 (2THF@EtP6, Fig. 3a and S21), one EtP6 molecule encapsulated two THF molecules in its cavity with C–H⋯O interactions, forming a 1 : 2 host–guest complex. Although 1D channels are observed, EtP6 adopts a slightly different conformation, caused by the presence of THF. Moreover, the PXRD pattern of EtP6β after adsorption of THF vapor matches well with that simulated from 2THF@EtP6, which is evidence for the structural transformation upon adsorption. In the crystal structure of pyrrolidine-loaded EtP6 (pyrrolidine@EtP6, Fig. 3b and S23), a 1 : 1 host–guest complex with pyrrolidine is found. Driven by C–H⋯π interactions and C–H⋯O hydrogen bonds formed by hydrogen atoms on pyrrolidine and oxygen atoms on EtP6, one pyrrolidine molecule is in the cavity of EtP6 with the nitrogen atom inside the cavity. The window-to-window packing mode of hexagonal EtP6 molecules in pyrrolidine@EtP6 contributes to the formation of honeycomb-like infinite edge-to-edge 1D channels, favorable for guest adsorption. Likewise, the PXRD result of EtP6β after adsorption of pyrrolidine is in line with the simulated pattern of pyrrolidine@EtP6, indicating that EtP6β transformed into pyrrolidine@EtP6 in the presence of pyrrolidine (Fig. S22 and S24).Open in a separate windowFig. 3Single crystal structures: (a) 2THF@EtP6; (b) pyrrolidine@EtP6.According to the adsorption ability and different crystal structures after adsorption of guest vapors, we wondered whether EtP5α or EtP6β could separate mixtures of THF and pyrrolidine. We first evaluated separation by EtP5α. GC analysis indicated that the adsorption ratios of THF and pyrrolidine were 65.7% and 34.3%, respectively, when EtP5α was exposed to 50 : 50 (v/v) pyrrolidine/THF mixture vapor (Fig. 4a and S25). Such adsorption was also illustrated by 1H NMR (Fig. S26). Although EtP5α showed a preference for THF, the selectivity is not satisfactory and cannot be applied to industrial separation. The less satisfactory selectivity may be ascribed to the similar crystal structures of EtP5 after adsorption of THF or pyrrolidine and insufficient strong stabilizing interactions. The PXRD pattern of EtP5α after adsorption of the 50 : 50 (v/v) pyrrolidine/THF mixture vapor exhibited minor differences compared with that simulated from either 2THF@EtP5 or pyrrolidine@EtP5, due to poor selectivity (Fig. 4b).Open in a separate windowFig. 4(a)Time-dependent solid–vapor adsorption plot for EtP5α in the presence of 50 : 50 (v/v) pyrrolidine/THF mixture vapor. (b) PXRD patterns of EtP5α: (I) original EtP5α; (II) after adsorption of THF vapor; (III) after adsorption of pyrrolidine vapor; (IV) after adsorption of pyrrolidine/THF mixture vapor; (V) simulated from the single crystal structure of pyrrolidine@EtP5α; (VI) simulated from the single crystal structure of 2THF@EtP5α. (c) Time-dependent solid–vapor adsorption plot for EtP6β in the presence of 50 : 50 (v/v) pyrrolidine/THF mixture vapor. (d) PXRD patterns of EtP6β: (I) original EtP6β; (II) after adsorption of THF vapor; (III) after adsorption of pyrrolidine vapor; (IV) after adsorption of pyrrolidine/THF mixture vapor; (V) simulated from the single crystal structure of pyrrolidine@EtP6β; (VI) simulated from the single crystal structure of 2THF@EtP6β.Nevertheless, selective separation of THF and pyrrolidine was achieved with EtP6β. As shown in Fig. 4c, time-dependent solid–vapor adsorption experiments for a 50 : 50 (v/v) pyrrolidine/THF mixture were conducted. Unlike the phenomenon in single-component adsorption experiments, uptake of pyrrolidine by EtP6β increased and reached the saturation point rapidly (less than 2 hours), while capture of THF was negligible. According to the NMR and GC results (Fig. S27 and S28), the purity of pyrrolidine was determined to be 99.9% after 2 hours of adsorption, which indicates the remarkable selectivity of EtP6β for pyrrolidine. The PXRD pattern of EtP6β after adsorption of the mixture was consistent with that from single-component adsorption, indicating the structural transformation in the crystalline state upon selective capture of pyrrolidine from the mixture. Although THF and pyrrolidine have similar molecular structures, their non-covalent interactions with EtP6 are different. We assume that the hydrogen bond between N–H and the oxygen atom on EtP6 stabilizes pyrrolidine and leads to such selectivity. More importantly, compared with previous adsorption processes using NACs reported by our group, the selective separation of pyrrolidine was completed rapidly. According to the GC results, the purity of pyrrolidine reached around 95% in the initial 15 min, while it usually takes hours for selective separations of other substrates using NACs. Increasing the adsorption time to 2 h improves the purity to over 99%. The rapid separation of pyrrolidine with high purity using EtP6β shows great potential in industrial applications.Apart from selectivity, recyclability is also an important parameter for an adsorbent. Consequently, recycling experiments were carried out by heating pyrrolidine@EtP6 under vacuum at 100 °C to remove adsorbed pyrrolidine. According to TGA and PXRD analysis, the recycled EtP6 solid maintained crystallinity and structural integrity that were the same as those of activated EtP6 crystals (Fig. S29 and S30). Besides, it is worth mentioning that the recycled EtP6 solids were still capable of separating mixtures of pyrrolidine and THF without loss of performance after being recycled five times (Fig. S31).In conclusion, we explored the separation of pyrrolidine/THF mixtures using NACs of EtP5 and EtP6. Pyrrolidine was purified using EtP6 from a 50 : 50 (v/v) pyrrolidine/THF mixture with a purity of 99.9%, but EtP5 exhibited selectivity towards THF. Moreover, the separation of pyrrolidine by EtP6 was extremely fast so that over 95% purity was determined within 15 min of adsorption. The rapid separation is unique among NAC-based separations. Single-crystal structures revealed that the selectivity depended on the stability of the new structures after adsorption of the guests and the non-covalent interactions in the host–guest complexes. PXRD patterns indicated that the structures of the host crystals changed into the host–guest complexes after adsorption. Additionally, the NACs of EtP6 exhibited excellent recyclability over at least five runs; this endows EtP6 with great potential as an alternative adsorbent for rapid purification of pyrrolidine that can be applied in practical industry. The fast separation with such simple NACs in this work also reveals that minor structural differences can cause significant changes in properties, which should provide perspectives on designs of adsorbents or substrates with specifically tailored binding sites.  相似文献   

17.
Covalent triazine-based frameworks (CTFs) have emerged as some of the most important materials for photocatalytic water splitting. However, development of CTF-based photocatalytic systems with non-platinum cocatalysts for highly efficient hydrogen evolution still remains a challenge. Herein, we demonstrated, for the first time, a one-step phosphidation strategy for simultaneously achieving phosphorus atom bonding with the benzene rings of CTFs and the anchoring of well-defined dicobalt phosphide (Co2P) nanocrystals (∼7 nm). The hydrogen evolution activities of CTFs were significantly enhanced under simulated solar-light (7.6 mmol h−1 g−1), more than 20 times higher than that of the CTF/Co2P composite. Both comparative experiments and in situ X-ray photoelectron spectroscopy reveal that the strong interfacial P–C bonding and the anchoring of the Co2P cocatalyst reverse the charge transfer direction from triazine to benzene rings, promote charge separation, and accelerate hydrogen evolution. Thus, the rational anchoring of transition-metal phosphides on conjugated polymers should be a promising approach for developing highly efficient photocatalysts for hydrogen evolution.

Reversing the electron transfer in a covalent triazine-based framework by Co2P anchoring achieved highly efficient photocatalytic hydrogen evolution from water splitting.

Photocatalytic water splitting into hydrogen fuels has been considered as a promising technique for converting solar energy into chemical energy.1–3 To achieve this target, it is necessary to design and construct photocatalysts with high solar-to-hydrogen (STH) conversion efficiency.4,5 Recently, covalent triazine-based frameworks (CTFs),6,7 as a new class of conjugated polymer materials, have attracted significant attention in the photocatalytic water splitting field owing to their visible-light response, organized architecture, adjustable pore-size, and controllable functionalization.8–12 However, owing to the high charge recombination and fewer active sites for the hydrogen evolution reaction (HER), the photocatalytic activities of pristine CTFs are very low. To overcome this drawback, noble-metal platinum (Pt) is generally required in CTF-based photocatalytic systems as the HER cocatalyst for promoting charge separation and catalyzing hydrogen generation.13–15 However, the high cost and scarcity of metallic Pt greatly limit the large-scale commercial applications. Thus, it is highly desirable to explore economical materials as noble-metal substitutes for achieving comparable or even superior hydrogen evolution activities.Recently, transition-metal phosphides (TMPs) have attracted intensive attention as HER cocatalysts for photocatalytic water splitting, owing to their unique structural and electronic properties.16–20 Up to now, a variety of semiconductor materials, including metal oxides,21,22 metal sulfides,23,24 g-C3N4,25,26 MOFs,27,28etc., decorated with TMPs have been extensively investigated, and the hydrogen generation activity in some reports is even higher than that of Pt cocatalysts. However, related studies about the decoration of TMP cocatalysts on CTF photocatalysts for highly efficient hydrogen generation have not been reported so far. Taking into account the molecular structure of CTFs, the nitrogen sites in the triazine frameworks could easily coordinate with transition metal ions to form nitrogen–metal interactions.29–31 In contrast, the coordination of TMPs with CTFs is relatively difficult owing to the high electronegativity of P sites which could draw electrons from metal atoms.17 Moreover, most reported TMPs were fabricated directly from precursor metal salts, oxides, etc.,32–35 and the resultant large-dimensions usually lead to very limited contact-interface with CTFs, or even a physical mixture form. Accordingly, the charge transfer between TMPs and CTFs is significantly restrained. Thus, effective anchoring of TMPs on CTFs for achieving highly efficient photocatalytic hydrogen evolution still remains a great challenge.Herein, a one-step phosphidation strategy has been developed to achieve phosphorus bonding with the benzene rings of CTFs and the anchoring of well-defined dicobalt phosphide (Co2P) nanocrystals (∼7 nm). The photocatalytic results clearly reveal that an excellent hydrogen evolution activity (7.6 mmol h−1 g−1) is achieved, which is much higher than that of the CTF/Co2P composite (0.37 mmol h−1 g−1). More detailed studies confirm that the phosphidation strategy could effectively facilitate the interfacial bonding between CTFs and Co2P nanocrystals. More importantly, the SI-XPS results clearly suggest that the charge transfer direction in CTFs is completely reversed, and the photo-generated electrons efficiently transferred from the triazine rings to P-bonded benzene rings, where the Co2P cocatalyst attracted electrons through the interfacial P–C bonds for efficient hydrogen evolution. To our knowledge, this is the first report on incorporating TMPs on a CTF photocatalyst for enhancing the hydrogen evolution activity. Fig. 1A shows the basic procedures for the preparation of CTF polymer anchored Co2P nanocrystals. Briefly, the CTFs with adsorbed cobalt ions were directly phosphatized by thermal decomposition of NaH2PO2 under an Ar atmosphere, which could simultaneously achieve P atom bonding with benzene rings and the anchoring of the Co2P cocatalyst (marked as P-CTF-Co2P). Fig. 1B shows the typical transmission electron microscopy (TEM) image of the obtained sample, clearly revealing that well-defined Co2P nanocrystals with an average diameter of 5∼9 nm were uniformly dispersed on the CTF surface. Furthermore, the high-resolution TEM (HR-TEM) image (Fig. 1C) of the formed nanocrystals exhibited two lattice fringes with d-spacing values of 0.20 and 0.21 nm, respectively, which could be well indexed to the (211) and (121) planes of orthorhombic Co2P crystals.36,37Fig. 1D shows the X-ray diffraction (XRD) pattern of the obtained P-CTF-Co2P sample. For comparison, the pristine CTF, and Co2P nanocrystals have also been studied. It can be clearly seen that for the P-CTF-Co2P sample, except for the two broad diffraction peaks at 7.3° and 26.1° attributed to the in-plane (100) facets and interlayer (001) stacking of CTFs,38,39 the other XRD peaks well matched with those of the Co2P crystals,40,41 confirming the successful incorporation of the Co2P cocatalyst in the CTFs. Furthermore, energy-dispersive X-ray (EDX) elemental mapping (Fig. 1E) clearly reveals the uniform distribution of C, N, Co and P elements in the whole detection region, further confirming the uniform dispersion of Co2P on CTF polymers.Open in a separate windowFig. 1(A) Basic procedures and the ideal structure scheme for fabricating P-CTF-Co2P catalysts; (B) TEM image and (C) HR-TEM image of the P-CTF-Co2P catalyst; (D) XRD patterns of CTFs, Co2P and P-CTF-Co2P catalysts; (E) EDX elemental mapping images of the P-CTF-Co2P catalyst.To further explore the P-bonding sites, high-resolution X-ray photoelectron spectroscopy (XPS) was performed on both P-CTF-Co2P and pristine CTFs. As shown in Fig. 2A, for pristine CTF samples, the C 1s peak could be well fitted into two peaks located at 284.8 and 286.8 eV, which could be assigned to C–C Created by potrace 1.16, written by Peter Selinger 2001-2019 C and N–C Created by potrace 1.16, written by Peter Selinger 2001-2019 N bonds,38,39,42 respectively. Notably, after the phosphidation treatment, a new peak at 285.3 eV attributed to P–C bonds was detected in the P-CTF-Co2P sample,43–45 and the C–C Created by potrace 1.16, written by Peter Selinger 2001-2019 C to N–C Created by potrace 1.16, written by Peter Selinger 2001-2019 N ratio was significantly decreased from 10.2 (pristine CTFs) to 7.1 (P-CTF-Co2P). In contrast, compared with the N 1s spectrum of the pristine CTF sample, no evident change of peak shape and intensity could be detected in the P-CTF-Co2P sample (Fig. 2B). These results clearly reveal that after the phosphidation treatment, P atoms should mainly bond with the carbon sites of the benzene rings in CTFs instead of the triazine rings.46,47 To further confirm this inference, XPS studies on phosphatized CTFs without Co2P nanocrystals (marked as P-CTFs) were also performed. It can be clearly seen from Fig. S9 that similar changes for the C 1s and N 1s peaks of P-CTF-Co2P (Fig. 2A and B) are observed in the P-CTF sample. Moreover, in both P-CTF-Co2P and P-CTF samples, an evident P 2p peak corresponding to P–C bonds could be observed (Fig. 2C),48 further indicating the bonding of P atoms with C atoms on the benzene rings in CTFs after the phosphidation treatment. However, note that the binding energy (BE) position of P–C peaks in P-CTF-Co2P (133.6 eV) is slightly higher than that of P-CTF (133.2 eV), which should be attributed to the anchoring with Co2P nanocrystals. In addition to the XPS investigations, solid-state cross-polarization magic angle spinning carbon-13 and phosphorus-31 nuclear magnetic resonance (13C-NMR and 31P-NMR) spectroscopy studies were further performed to explore the bonding sites of phosphorus atoms in the P-CTF (Fig. S10). The Fourier-transformed infrared (FTIR) spectrum was also employed (Fig. 2D). It can be clearly observed that the typical FTIR peaks of the triazine frameworks and the stretching vibrations of carbon–nitrogen (C–N) (1521 cm−1 and 1354 cm−1) in P-CTF-Co2P are consistent with those of pristine CTFs.38,39 However, the peaks at 1672 cm−1 and 1014 cm−1 corresponding to the stretching vibrations of C Created by potrace 1.16, written by Peter Selinger 2001-2019 C and C–H in benzene rings were evidently decreased in P-CTF-Co2P compared with those of pristine CTFs.49,50 Combining the results of XPS, NMR, and FTIR, it can be concluded that in the obtained P-CTF-Co2P sample, the carbon sites of the benzene rings in the CTFs were partially bonded with P-atoms, which should anchor Co2P through P–Co bonding.Open in a separate windowFig. 2High-resolution XPS spectra of (A) C 1, (B) N 1s and (C) P 2p in CTF, P-CTF and P-CTF-Co2P samples (the insets show the molecular structures); (D) Fourier transform infrared spectroscopy (FT-IR) of CTF and P-CTF-Co2P samples.Furthermore, the hydrogen evolution activities of the P-CTF-Co2P (2 wt% Co2P) sample were evaluated under simulated solar light irradiation. For comparison, the CTF/Co2P composite and pristine CTFs were also measured under the same conditions. As shown in Fig. 3A, the P-CTF-Co2P sample exhibits a much higher H2 evolution activity (15.2 mmol g−1) than the CTF/Co2P composite (0.7 mmol g−1) at 2 h, while no evident hydrogen generation could be detected in the pristine CTF sample. To further explore the crucial roles of interfacial P-bonding in CTFs and the anchoring of the Co2P cocatalyst, the hydrogen evolution rates of various samples were calculated and their comparison is shown in Fig. 3B. Obviously, compared with the excellent activity of P-CTF-Co2P (7.6 mmol h−1 g−1), CTF/Co2P and P-CTF/Co2P only exhibit very low hydrogen evolution rates of 0.37 and 0.38 mmol h−1 g−1, respectively, while no evident activity could be detected for CTFs, Co2P, and P-CTF samples. The above results clearly confirm that the significant enhancement of the hydrogen evolution activity of the P-CTF-Co2P sample should be mainly attributed to the P-bonding and the anchoring of the Co2P cocatalyst. Furthermore, the photocatalytic performances of CTFs decorated with Pt nanoparticles were also studied (the inset of Fig. 3B, Fig. S15), as Pt is generally recognized as the most active cocatalyst for hydrogen generation. Amazingly, the hydrogen production rates of the Pt cocatalyst at different amounts (1∼5 wt%) were all lower than that of the P-CTF-Co2P sample, indicating that the rational bonding of the Co2P cocatalyst on CTFs should be a promising strategy for achieving highly efficient photocatalytic hydrogen evolution.Open in a separate windowFig. 3(A) Photocatalytic H2 evolution tests of the as-prepared photocatalysts; (B) comparative presentation of the hydrogen evolution rates; (C) ultraviolet-visible diffuse reflectance spectrum of the P-CTF-Co2P photocatalyst (red solid line) and wavelength-dependent hydrogen production activities of the P-CTF-Co2P photocatalyst within 2 h; light source: a 300 W Xe lamp equipped with various cut-off filters. (D) The cycling photocatalytic tests of P-CTF-Co2P and the CTF/Co2P composite; (E) It curves of various catalysts at −0.3 V (vs. SCE); (F) electrochemical impedance spectroscopy under light of the various catalysts. Measurements were conducted in 0.2 mol L−1 Na2SO4 electrolyte solution under AM 1.5 G illumination.Furthermore, the relationship between wavelength and the hydrogen evolution activity of the P-CTF-Co2P photocatalyst was studied and is shown in Fig. 3C. With increasing the wavelength from full-spectrum light irradiation (λ > 300) to 500 nm, the hydrogen evolution rate significantly decreased from 7.6 to 0.06 mmol h−1 g−1, which is generally consistent with their absorption spectrum. The highest value of the apparent quantum yields (AQYs) attained is 31.8% at 365 nm. Moreover, the photocatalytic stability and durability of P-CTF-Co2P and CTF/Co2P were examined by cycling experiments. As shown in Fig. 3D, the CTF/Co2P composite demonstrated relatively poor stability, and there is no evident activity after only one cycling experiment, resulting from the incompact combination between CTFs and Co2P (Fig. S5). In contrast, the P-CTF-Co2P sample exhibits relatively high stability, and no evident inactivation was detected during the whole test. These results further confirm that in addition to enhancing activities, the interfacial P–C bonds and the anchoring of the Co2P cocatalyst could also effectively promote the hydrogen evolution stability. To further confirm the efficient charge separation and electron transfer in P-CTF-Co2P, photoelectrochemical (PEC) tests were performed. As shown in Fig. 3E, the amperometric It curves clearly reveal that P-CTF-Co2P exhibits higher photocurrent density than pristine CTFs and CTF/Co2P samples, confirming its more efficient charge separation capability. Moreover, electrochemical impedance spectroscopy (EIS) was also performed to explore the interfacial charge transfer process.51,52 As shown in Fig. 3F, the P-CTF-Co2P sample with the smallest arc radius revealed remarkably increased interfacial charge transport efficiency. These PEC results clearly demonstrate the highly efficient photogenerated charge separation and transfer in the P-CTF-Co2P photocatalyst, which are highly consistent with the above photocatalytic hydrogen evolution results.The photo-induced charge separation and transfer in the excited state are crucial processes for determining the photocatalytic activity. Herein, we have demonstrated an in situ irradiation X-ray photoelectron spectroscopy (SI-XPS) technique for exploring the intrinsic charge separation and transfer mechanisms between Co2P and CTFs. As shown in Fig. S20 and S21, no evident binding energy (BE) shift could be observed in both pristine CTF and Co2P samples, confirming their relatively poor charge separation capability. Amazingly, for the P-CTF-Co2P sample (Fig. 4), distinct variations for C 1s, N 1s, P 2p, and Co 2p peaks were detected in the excited state. More specifically, the C 1s and N 1s peaks in the CTFs shifted towards the high BE region by 0.3 eV accompanied by the broadening of peak shape (Fig. 4B and C), while the P 2p and Co 2p peaks of the Co2P cocatalyst shifted towards the low BE direction by 0.3 and 0.4 eV (Fig. 4D and E), respectively.Interestingly, as shown in Fig. 4D, the P 2p peaks attributed to interfacial P–C bonds exhibited no evident BE change under dark and light irradiation. On the basis of the above SI-XPS results, it can be concluded that under light irradiation, the photo-excited electrons effectively transferred from CTFs to Co2P through the P-bonding sites as well as the interfacial P–C bonds, leading to electron enrichment in Co2P nanocrystals and hole enrichment in CTFs. To further confirm the crucial roles of interfacial P-bonding in CTFs and the anchoring of the Co2P cocatalyst in promoting charge separation, SI-XPS studies for the CTF/Co2P composite were also conducted. As shown in Fig. S22, no new peak or shape change of C 1s, N 1s, P 2p, and Co 2p peaks could be detected in CTF/Co2P compared with pristine CTFs and Co2P in the ground state, indicating no interfacial bonding in the CTF/Co2P composite. Furthermore, in the excited state, no evident BE shifts and shape change could be observed, indicating the poor charge-separation capability of CTF/Co2P, which is highly consistent with its fairly low hydrogen evolution activity. These SI-XPS results further confirm the crucial roles of interfacial P-bonding in CTFs and the anchoring of the Co2P cocatalyst in efficiently promoting charge separation and enhancing the photocatalytic activity.Open in a separate windowFig. 4(A) Schematic illustration of the SI-XPS technique for direct observation of electron transfer in the excited state. The (B) C 1s, (C) N 1s, (D) P 2p and (E) Co 2p of SI-XPS spectra in the P-CTF-Co2P sample tested under dark and light illumination.On the basis of the above results, it can be concluded that the P-bonding and Co2P-cocatalyst anchoring should reverse the photo-induced electron transfer direction from the triazine to benzene rings in CTFs. To further confirm this speculation, the photocatalytic behavior of various CTF-based photocatalysts was studied and their electron transfer directions in excitation states have been proposed (Fig. 5A and B). First, when Co2+ ions were further bonded with the N sites of the triazine rings in the P-CTF-Co2P sample (marked as P-CTF-Co2P-Co), the hydrogen evolution activity remarkably increased from 7.6 up to 8.4 mmol h−1 g−1. This result clearly reveals that the Co–N bonding in P-CTF-Co2P could further promote charge separation due to the effective hole trapping on Co sites. In contrast, when the Pt cocatalyst was decorated on the P-CTF-Co2P sample (marked as P-CTF-Co2P–Pt) to form Pt–N coordination, the photocatalytic activity significantly decreased to 4.6 mmol h−1 g−1, mainly resulting from the electron transfer competition between Pt and Co2P. Furthermore, the decoration of the single Pt cocatalyst on CTFs promoted the photo-induced electron transfer from the benzene rings to triazine rings, and the photocatalytic activity achieved was up to 3.8 mmol h−1 g−1. Furthermore, in situ FTIR spectroscopy was performed for exploring the intermediate products in the photocatalytic process of the P-CTF-Co2P sample with co-adsorption of H2O vapor (Fig. S27), which also demonstrated the direction of electron transfer combined with the in situ XPS results. Accordingly, a possible mechanism has been proposed for clarifying the hydrogen evolution activities of the P-CTF-Co2P photocatalyst (Fig. 5B). Owing to the P-bonding and the anchoring of Co2P nanocrystals, the photo-generated electrons could effectively transfer from the triazine rings to benzene rings, where the electrons are trapped by the Co2P cocatalyst through the interfacial P–C bonds for the hydrogen evolution reaction. Simultaneously, the holes left behind in the CTFs participated in oxidation reactions. More importantly, this mechanism may provide a new insight for understanding the crucial roles of interfacial P-bonding in CTFs and their bonding with Co2P in promoting the charge separation for efficient hydrogen evolution.Open in a separate windowFig. 5(A) Photocatalytic hydrogen evolution activities of various samples and the scheme of electron-transfer direction in the excited state. (B) The ideal structural illustration of the interfacial bonding and charge transfer of the P-CTF-Co2P photocatalyst for hydrogen evolution.  相似文献   

18.
Correction for ‘Click activated protodrugs against cancer increase the therapeutic potential of chemotherapy through local capture and activation’ by Kui Wu et al., Chem. Sci., 2021, 12, 1259–1271, DOI: 10.1039/D0SC06099B.

The authors regret that the reference to the bond-breaking bioorthogonal chemistry, termed ‘click-to-release’ was omitted from the original article. In addition, we would like to include a reference describing the synthesis of compound 1, which is an intermediate to the protodrugs described in the original article. These references are listed below as ref. 1 and 2.The Royal Society of Chemistry apologizes for these errors and any consequent inconvenience to authors and readers.  相似文献   

19.
Heteroatom-containing degradable polymers have strong potential as sustainable replacements for petrochemically derived materials. However, to accelerate and broaden their uptake greater structural diversity and new synthetic methodologies are required. Here we report a sequence selective ring-opening terpolymerisation (ROTERP), in which three monomers (A, B, C) are selectively enchained into an (ABA′C)n sequence by a simple lithium catalyst. Degradable poly(ester-alt-ester-alt-trithiocarbonate)s are obtained in a Mn range from 2.35 to 111.20 kDa which are not easily accessible via other polymerisation methodologies. The choice of alkali metal is key to achieve high activity and to control the terpolymer sequence. ROTERP is mechanistically compatible with ring-opening polymerisation (ROP) allowing switchable catalysis for blockpolymer synthesis. The ROTERP demonstrated in this study could be the first example of an entirely new family of sequence selective terpolymerisations.

A sequence selective ring-opening terpolymerisation of epoxides with CS2 and phthalic thioanhydride yielding poly(ester-alt-ester-alt-trithiocarbonates) is reported.

Synthetic polymers are now more in demand than ever before and looking at their annually increasing production, a polymer free society is at best a vague memory rather than a vision for the future.1 As most commodity polymers are based on chemically inert aliphatic –C–C– backbones, most polymer waste shows unappreciable degradation with respect to the polymer''s time in application.2,3 In answer to the ever-increasing amount of plastic pollution, much effort focuses on the exploration of new heteroatom containing polymers which, because of their more polar bonds making up the backbone, are more susceptible to degradation via physical, chemical and biochemical pathways and even facilitate new chemical recycling methods.4–8 Besides, there is also a constant demand for entirely new materials to enable technological innovation making new methodologies to synthesise heteroatom containing polymers necessary.Arguably one of the most popular methods to make such polymers is the ring-opening polymerisation (ROP) of a heterocycle A.9–11 These polymerise under release of their associated ring strain energy to make polymers (A)n such as poly(thio)ethers, poly(thio)ester and poly(thio)carbonates, only to name a few. Early on it has been realised that in some cases the ROP of three or four-membered heterocycles A can be coupled to the insertion of typically heteroallenes or cyclic anhydrides B to generate alternating copolymers (AB)n.12–14 The underlying requirements for (AB)n polymerisations are a combination of kinetic factors, i.e. monomer A inserting orders of magnitude faster into the active catalyst growing chain-bond than monomer B, and chemoselectivities, i.e. insertion of monomer A resulting in type A active catalyst growing chain-bond that does not show appreciable reactivity with A but only with B (and vice versa). Prominent examples of this alternating (AB)n ring-opening copolymerisation (ROCOP) include CO2/epoxide ROCOP yielding polycarbonates and cyclic anhydride/epoxide ROCOP yielding polyesters.15–17 Sulfur analogous are also accessible such as polythiocarbonate from CS2/(epoxide or thiirane) and polythioesters from thioanhydride/(epoxide or thiirane) ROCOP.18–31 Such sulfur rich polymers are attractive high-refractive index materials that can show improved crystallinity and degradability over their all-oxygen analogues and in some cases enable chemical polymer to monomer recycling.32–39Most relevant to this study is a report by Werner and coworkers on lithium alkoxide catalysed CS2/epoxide ROCOP yielding poly(monothio-alt-trithiocarbonate)s featuring R–O–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–O–R and R–S–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–S–R links (Fig. 1).40,41 Such a polymer sequence is unexpected, as the formal product of alternating insertion would be a poly(dithiocarbonate) with R–O–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–S–R links. Furthermore, the polymer shows an unusual “head-to-head-alt-tail-to-tail” selectivity meaning that the R–O–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–O–R links sit adjacent to the tertiary CHR3 positions of the ring opened epoxide (i.e. “head” position) and that the R–S–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–S–R links sit adjacent to the secondary CH2R2 position of the ring opened epoxide (i.e. “tail” position). This sequence let the authors postulate a mechanism involving tail-selective epoxide ring-opening by a dithiocarbonate chain end which is formed by CS2 insertion into an alkoxide intermediate. The resulting alkoxide intermediate was proposed to backbite into the adjacent dithiocarbonate link which after a rearrangement process resulted in an O/S exchange of the chain end. The rearrangement transforms the alkoxide into a thiolate chain end and the adjacent dithiocarbonate R–O–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–S–R into a monothiocarbonate R–O–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–O–R. CS2 insertion of the thiolate generates a trithiocarbonate R–S–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–S–R which after epoxide insertion regenerates the alkoxides. In contrast to alkoxides sitting adjacent to R–O–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–S–R links, alkoxides next to R–S–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–S–R links were not proposed to undergo backbiting and O/S exchange. Importantly the initial regioselectivity of the epoxide ring-opening was preserved throughout the rearrangement which explained the “head-to-head-alt-tail-to-tail” selectivity. Interestingly lithium appeared to be crucial as other alkali metal alkoxides failed to catalyse this ROCOP, while more sophisticated catalysts result in much more uncontrolled polymer sequences. Hence it appears that the Li controls which alkoxide intermediate precisely undergoes O/S exchange and to which degree this rearrangement occurs, but the reasons behind the special role of Li remains to be explored. Relatedly the ROCOP of thioanhydrides with epoxides has also been reported and similar exchange processes have been proposed as side reactions.23 Although not directly proven, this mechanism seemed reasonable and let us hypothesise that lithium catalysts could grant a general access to control the O/S exchange process in ROCOP and even in the polymerisation of ternary monomer mixtures. Furthermore, we reasoned that the two distinct chain ends formed via O/S exchange, i.e. alkoxide and thiolate (type A vide supra), could enable discrimination between two different type B monomers and enable sequence selective terpolymerisations. It should be noted that reports exist in which mixtures of A, B and C (e.g. epoxide A, cyclic anhydride B and CO2 C) either yield random terpolymers, (AB)n-ran-(AC)m poly(esters-ran-carbonate), or block-terpolymers, (AB)n-b-(AC)m, polyester-b-polycarbonate. In this case the monomer sequence depends on catalyst selection and reaction conditions, but sequence selective terpolymers, e.g. (ABC)n or (ABAC)n, are unknown.42–48 The hypothesis of O/S exchange here led us to discover a new type of polymerisation, sequence selective ring-opening terpolymerisation (ROTERP) that we report in this contribution. ROTERP produces poly(ester-alt-ester-alt-trithiocarbonates), i.e. (ABA′C)n sequences, from a mixture of the monosubstituted epoxides propylene oxide (PO) or butylene oxide (BO) A, phthalic thioanhydride (PTA) B and CS2 C, employing simple lithium salts as the catalyst. Furthermore, model reactions proof the previously postulated O/S rearrangement that enable ROTERP and elucidate the role of the lithium catalyst. Finally, we employed ROTERP in so-called switchable catalysis, in which the onset of ROTERP stops the occurrence of ROP, for the construction of blockpolymers.Open in a separate windowFig. 1(Top) CS2/epoxide ROCOP and postulated mechanism involving a central O/S exchange reaction, (middle) phthalic thioanhydride (PTA)/epoxide ROCOP and (bottom) PTA/CS2/epoxide ROTERP reported in this study. R = Me, Et; [Rn] = polymer chain.ROTERP of mixtures of PTA, PO and CS2 at loadings typically employed in ROCOP catalysis with lithium hexamethyldisilazide (LiHDMS) or lithium benzyloxide (LiOBn) as the catalyst at loadings in the range of 1 eq. LiX : (6.25–500 eq.) PTA : (31.25–2500 eq.) PO : (62.5–5000 eq.) CS2 yield polymeric materials in 95% selectivity at 80 °C (see 40,47 Spectroscopic analysis of the isolated polymer reveals surprisingly clean NMR spectra given the potential for statistical terpolymerisation of these three monomers. The 1H NMR spectrum (Fig. 2) shows two main aryl resonances for a symmetrically substituted terephthalate unit from ring opened PTA (δ = 7.72 and 7.56 ppm) as well as one main resonance for the CHMe (δ = 5.42 ppm; head position of the ring-opened epoxide) and CH2 (δ = 3.90–3.40 ppm; tail position of the ring opened epoxide) groups respectively stemming from the ring opened PO. Correspondingly the 13C{1H} NMR spectrum (ESI Fig. S1) reveals the almost exclusive presence of trithiocarbonate R–S–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–S–R (δ = 222.8 ppm) and arylester R–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 O)–O–R (δ ∼166.6 ppm) resonances (94–98%) alongside minor thioester R–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 O)–S–R (δ = 192.7 ppm) resonances (2–6%). 1H and 2D NMR spectra (ESI Fig. S3) show that trithiocarbonate units are positioned adjacent to CH2 groups while arylesters are connected to the tertiary CHMe groups. The spectra remain unchanged after multiple precipitations from DCM : MeOH or THF : pentane and DOSY NMR shows that all 1H NMR resonances diffuse at the same rate confirming that these are part of the same species. Furthermore, no other type of thiocarbonate R–(O/S)–C( Created by potrace 1.16, written by Peter Selinger 2001-2019 O/S)–(O/S)–R are part of the polymer. Linkage identity was further substantiated by the ATR-IR spectrum (ESI Fig. S5) of these polymers showing an arylester C Created by potrace 1.16, written by Peter Selinger 2001-2019 O stretch at Created by potrace 1.16, written by Peter Selinger 2001-2019 = 1716 cm−1 as well as a thiocarbonate C Created by potrace 1.16, written by Peter Selinger 2001-2019 S stretch at Created by potrace 1.16, written by Peter Selinger 2001-2019 = 1062 cm−1 (ESI Fig. S5). Accordingly, the polymers are obtained as yellow solids due the presence of the C Created by potrace 1.16, written by Peter Selinger 2001-2019 S chromophore (λabs = 435 nm, ESI Fig. S6). MALDI-TOF analysis unfortunately only led to decomposition of the materials and no signals could be identified as previously reported for sulfur-rich polymers.26,38 Nevertheless 1H NMR allows some conclusions regarding the topology as both for LiHMDS or LiOBn initiation, resonances for the HMDS and OBn groups can be identified to be part of the purified polymers. Insertion of alkalimetal alkoxides and amides into CS2 yielding alkali dithiocarbonates and dithiocarbamates have been previously reported.40,79 This makes initiation via CS2 insertion likely which defines one end of the polymer and hence indicates that chains are linear rather than cyclic. As ROTERP is followed by CS2/epoxide coupling once all PTA is consumed (vide infra) we infer that chains are terminated by CS2 because heteroallene insertion products are generally established to be the resting states of heteroallene/heterocycle coupling reactions.12Data showing PTA/CS2/epoxide ROTERP under different conditions
RunLiXf : PTA : (PO/*BO) : CS2aTime [min]PTA conv.Polym. select.bLinkage select.c M n [kDa] (Đ)d M n,theo [kDa]
#11 : 6.25 : 31.25 : 62.5<1 min>99%95%98%2.35 (1.44)2.41
#21 : 12.5 : 62.5 : 125<1 min>99%95%98%5.11 (1.41)4.81
#31 : 25 : 125 : 2501 min>99%95%98%8.90 (1.53)9.46
#41 : 100 : 500 : 100015 min>99%95%95%24.46 (1.47)37.34
#5e1 : 300 : 1500 : 300060 min98%95%95%55.05 (1.60)111.70
#6e1 : 500 : 2500 : 500016 h93%95%94%111.20 (1.76)186.06
#71 : 100 : 500* : 25015 min>99%95%91%23.45 (1.54)38.67
#81 : 100 : 500* : 50030 min>99%95%92%24.86 (1.67)38.67
#91 : 100 : 500* : 1000120 min>99%95%96%22.90 (1.55)38.67
#101 : 100 : 500* : 1500120 min>99%95%97%24.16 (1.56)38.67
#111 : 300 : 500 : 1000120 min90%95%91%55.05(1.55)104.20
#121g : 100 : 500* : 1000120 min22%95%77%4.20 (1.24)8.51
#131h : 100 : 500* : 1000120 min
#141 : 0 : 500* : 100018 h0%
#151 : 100 : 500* : 036 h76%99%40%7.17 (1.47)18.03
#16i1 : 100 : 500* : 100030 min>99%95%96%22.69 (1.55)38.67
#17j1 : 100 : 500* : 100010 min>99%95%96%23.73 (1.60)38.67
Open in a separate windowaCopolymerisation at T = 80 °C.bPolymer selectivity, determined by comparison of the relative integrals, in the normalised 1H NMR spectrum (CDCl3, 25 °C, 400 MHz), of tertiary CH resonances due to polymer and cyclic dithiocarbonate c5c at 20–80% PTA consumption.cLinkage selectivity, determined by comparison of the relative integrals, in the normalised the 1H NMR spectrum (CDCl3, 25 °C, 400 MHz) of resonances due to ester and trithiocarbonate linkages relative to ester, trithiocarbonate and thioester links (for #9 proportion of terephthalate and dithioterephthalate links).dDetermined by SEC (size exclusion chromatography) measurements conducted in THF, using narrow MW polystyrene standards to calibrate the instrument.eLonger reaction time was chosen due to high viscosity of the reaction mixture.fX = HMDS or OBn from in situ reaction of LiHMDS with 1 eq. BnOH.gNaHMDS with 1 eq. BnOH was employed as the catalyst.hKHMDS with 1 eq. BnOH was employed as the catalyst.i T = 100 °C.j T = 120 °C.Open in a separate windowFig. 2(Top left) PTA/CS2/PO ROTERP reaction scheme, X = HMDS, OBn; (top right) SEC trace corresponding to ) rendering it a useful methodology for future material synthesis.49 Aliquots removed at regular time intervals show a linear increase of molecular masses with PTA conversion with slightly increasing dispersity (ESI Fig. S27) which points towards some transesterification processes occurring alongside propagation, and this is further indicated by the presence minor CH2-ester resonances.50–52 Aliquot analysis by 1H NMR (ESI Fig. S23) shows uniformly growing polymer resonances forming in the reaction mixtures equivalent to those observed for the isolated polymer after full PTA consumption. This indicates poly(ester-alt-ester-alt-trithiocarbonate) formation throughout the reaction and no change of the respective link resonance ratios as ROTERP progresses pointing towards selective monomer enchainment rather than linkage formation through transesterification like processes. Taken together the results indicate a poly(ester-alt-ester-alt-trithiocarbonate) sequence as conveyed in Fig. 3 featuring a head-to-head connected terephthalate links and tail-to-tail connected trithiocarbonate links in alternation which is reminiscent of the results described by Werner and coworkers (vide supra).Open in a separate windowFig. 3Reaction products and postulated ROTERP reaction mechanism, [Rn] = polymer chain.The related regiochemistry in addition to the fact that a similar lithium catalyst generates alternating oxygen and sulfur enriched links let us hypothesize that the ROTERP process possesses mechanistic similarities to the ROCOP process reported by Werner and this led us to propose the propagation mechanism shown in Fig. 3. Here a thiocarboxylate intermediate TC, generated from alkoxide Alk insertion into PTA in step (i), inserts into the epoxide to form a thioester appended alkoxide Alk* in step (ii). This alkoxide then rearranges in an O/S exchange process into an ester appended thiolate T in step (iii). CS2 insertion by T forms a trithiocarbonate intermediate TTC, which inserts into BO in step (v) to regenerate Alk. This propagation results in a (ABA′C)n sequence with one link that derives from a ring opened epoxide A and another link from a ring opened epoxide following isomerisation (akin to a virtual thiirane) which we decided to term A′. In line with our mechanistic hypothesis, we believe that the erroneous thioester linkages are formed through incomplete O/S exchange and PTA insertion from Alk* or due to PTA insertion from the thiolate intermediate T as also shown in Fig. 3. The cyclic byproduct c5c is proposed to be formed via backbiting from Alk into the adjacent trithiocarbonate link where c5c elimination occurs over O/S exchange.To explore how thioester errors are formed, we conducted a series of terpolymerisation experiments in which we changed the PTA : CS2 loading from 100 : 1500 eq, to 100 : 250 eq. (versus 1 LiHMDS eq. and 500 BO eq.) which results in an effective concentration increase of PTA while decreasing the CS2 concentration. This results in a gradual increase of thioester links from 3% to 9% (). Increasing the PTA vs. CS2 loading from 100 : 1000 eq. to 300 : 1000 eq. likewise results in an increase in thioester links from 4 to 9%. Our results indicate kinetic competition between O/S exchange versus PTA insertion from Alk* and CS2versus PTA insertion from T. Furthermore, we find that the amount of thioester links remains constant when increasing the reaction temperature from 80 °C to 100 °C to 120 °C (). ROCOP between PTA and BO is also catalysed by LiOBn, which gave further insight into the ROTERP process. The produced polymers are colourless poly[(thio)ester]s featuring characteristic ester (δ = 165.8–167.5 ppm, Created by potrace 1.16, written by Peter Selinger 2001-2019 = 1725 cm−1) and thioester (δ = 192.0–193.3 ppm, Created by potrace 1.16, written by Peter Selinger 2001-2019 = 1670 cm−1) signals in NMR and IR (ESI Fig. S17–S21). Again 1H–13C HMBC NMR spectroscopy reveals that thioesters sit adjacent to the secondary CH2R2 tail position of the ring opened BO while esters sit adjacent to the tertiary CHR3 head positions but in contrast to the ROTERP case no long-range order can be observed. The formed polymer features 60% monothioteraphthalate (δ(13C) = 192.1 and 166.6 ppm), 20% dithioterephthalate (δ(13C) = 192.8 ppm) and 20% terephthalate links (δ(13C) = 165.9 ppm). For this ROCOP we also propose propagation via alternating enchainment of PTA and BO alongside O/S exchange at the chain end. Note that if O/S exchange was quantitative (or completely absent) in PTA/BO ROCOP one would only observe the formation of monothioterephthalate links. The formation of dithioterephthalate and terephthalate links alongside monothioterephthalate links however necessitate incomplete O/S exchange and insertion of lithiumthiolates alike T into PTA (Fig. 4). This makes it likely that both these pathways also occur in ROTERP causing the formation of thioester errors (Fig. 3). Furthermore, Li catalysed PTA/BO ROCOP is also significantly slower than ROTERP (TOF(ROTERP) > 100 h−1 and TOF(ROCOP) = 2 h−1, see Open in a separate windowFig. 4(Left) Selected region of the 1H–13C HMBC NMR spectrum (CDCl3, 25 °C, 500 MHz) of the polymer corresponding to Fig. 3) in which c5c elimination is favoured over O/S exchange. Accordingly attempted ROCOP between CS2 and BO at 80 °C exclusively yields c5c and no polymer (). We observe the same reactivity after full PTA consumption in ROTERP where the lithium catalyst switches from terpolymer production to c5c formation (ESI Fig. S23) and these results confirm that c5c is formed from backbiting reactions if PTA insertion does not occur. However, backbiting onto the trithiocarbonate links appears to be disfavoured in general as only small amounts of c5c (5% of the product mixture) are formed. In a related report on CS2 ROCOP, trithiocarbonate links have been observed to be the thermodynamic product of O/S scrambling suggesting that O/S exchange pathways that involve trithiocarbonates are thermodynamically unfavourable. The stability of the trithiocarbonate link could originate from resonance effects of the π-system which would also result in energetically less accessible π*-orbitals towards nucleophilic attack through backbiting.27,53Clearly the O/S exchange reaction is crucial for the occurrence of ROTERP. To verify and further explore this isomerisation step, we synthesised a model intermediate for TC namely MTC from the stoichiometric reaction of LiOtBu with PTA (Fig. 5) which instantaneously reacts in THF at room temperature similar to propagation step (i) in Fig. 3 (ESI Section S5). To obtain structural insight we crystallised MTC from THF which is serving as a model donor in place of epoxides. Intriguingly single crystal X-ray analysis shows the formation of a dimer where two lithium thiocarboxylates come together to form a central Li2O2 motif via coordination of the Ar-C( Created by potrace 1.16, written by Peter Selinger 2001-2019 S)–O oxygen and coordinative saturation with two THF molecules per Li. The bimetallic nature of MTC is interesting in light of recent developments in ROCOP showing that multimetallic complexes are particularly active in this catalysis and the same could be true for ROTERP given the strong tendency for lithium salts to form aggregates in solution.54–60 Furthermore the sulphur centres remain uncoordinated by Li, presumably due to its comparatively high oxophilicity in the series of alkali metals.61 Hence the sulphur centres are sterically unencumbered which might aid propagation through nucleophilic attack by those. The yellow (C Created by potrace 1.16, written by Peter Selinger 2001-2019 S) chromophore is maintained in THF solution (λAbs = 350–430 nm, ESI Fig. S37), however upon addition of excess BO gradual discolouration over the course of five minutes occurs. NMR analysis shows exclusive formation of ester containing products with no thioester resonances present (ESI Fig. S34). ESI-MS identifies the reaction products as phthalic diester appended thioethers (ESI Fig. S36). Our observations can be rationalised by insertion of MTC into BO via nucleophilic attack of the sulfur centre like step (ii) followed by O/S isomerisation as in step (iii) and consecutive insertion of the formed lithium thiolate into BO. The observed reactivity not only supports the mechanistic hypothesis outlined in Fig. 3 but also shows that lithium thiolates insert into BO which explains the formation of significant thioether links in PTA/BO ROCOP in absence of CS2. When MTC was reacted with substoichiometric (0.5 eq.) amounts BO to avoid thioether formation we observe clean formation of butylene thiirane and the corresponding carboxylate (Fig. 5 and ESI Fig. S32). This reactivity also confirms step (ii) and (iii), whereas now the thiolate intermediate reacts under intramolecular nucleophilic substitution to form a thiirane and eliminates the adjacent ester link as a carboxylate. Hence, we suggest that free thiolate chain ends appended to ester groups such as T are short living intermediates during ROTERP and the fact that no thiirane is observed during ROTERP also supports this. A different outcome is observed when excess CS2 (10 eq.) is present during the reaction of 1 eq. BO with MTC (ESI Fig. S38 and S39). We again only observe ester and no thioester containing products but also observe the initiation of CS2/BO ROCOP forming scrambled polythiocarbonate alongside c5c leaving 85% MTC unreacted. Hence the propagation steps that don''t involve MTC appear to be faster than (ii) which makes this the presumably slowest propagation step of ROTERP (note that PTA reacts on the timescale of seconds with LiOtBu while MTC insertion into BO occurs on the timescale of minutes to hours and that thiolates were found to be unstable towards thiirane elimination which also supports this notion). Furthermore, as we always observe quantitative O/S exchange, we suggest that this process is thermodynamically favoured and errors from incomplete O/S exchange during ROTERP are kinetic in origin.Open in a separate windowFig. 5Mechanistic experiments with selected regions of the 1H NMR spectra (CDCl3, 25 °C, 400 MHz) supporting O/S exchange reaction and solid-state structure of MTC, hydrogen atoms omitted for clarity; white = C; blue = O, yellow = S, purple = alkali metal.As outlined above the CS2 ROCOP literature shows that only lithium catalysts can control the O/S exchange process. To explore this for ROTERP we conducted terpolymerisation with NaOBn and KOBn in place of LiOBn (Fig. 6 and ESI Fig. S16). Employing KOBn results in no polymerisation at all. Intrigued by this striking difference in activity and selectivity for different alkali metals we conducted the analogous model experiments as outlined in the previous section with Na and K. Reaction of NaOtBu and KOtBu with PTA in THF at room temperature results in quantitative PTA ring-opening within seconds yielding the Na derivative MTCNa and the K derivative MTCK. Both crystallise as extended networks as can be seen in Fig. 6. In contrast to the lithium derivative MTC, coordination of the thiocarboxylate sulfur centre as well as the adjacent ester carbonyl oxygen centre is also observed in the solid-state structures of MTCNa and MTCK. While the alkali metal is four-coordinate in MTC as usually observed for Li, Na and K in MTCNa and MTCK are five and six coordinate. Although the precise structure in solution of these model intermediate remains to be determined, we still believe that because all structures were obtained under identical conditions (i.e. from THF/Pentane mixtures) our results highlight the greater tendency for the softer and larger alkali metals coordinated to the sulfur centres as well as the functionalities of the adjacent polymer chain. Hence, we propose that the rigid coordination sphere in addition to the high oxophilicity of lithium are responsible for its activity in ROTERP. We next reacted MTCNa with 1 eq. BO in presence of excess CS2 in THF (Fig. 6, ESI Fig. S42). In contrast to MTC which reacts within minutes with BO at room temperature, MTCNa reacts with BO on the timescale of hours, and this reflects the reduced activity of Na in ROTERP compared to Li. NMR analysis of the product mixture reveals the formation of c5c (δ(13C) = 210.8 ppm) and a diester appended anionic trithiocarbonate (δ(13C) = 243.6 ppm) as the main reaction products alongside unconsumed MTCNa.62,63 As for Li, we only observe ester and no thioesters containing products also pointing towards quantitative O/S exchange for Na. In ROTERP however Na produced significantly more thioester links from incomplete O/S exchange and PTA insertion into thiolate chain ends. Hence our findings indicate that O/S exchange versus PTA insertion selectivity could be kinetically controlled by the metal catalyst. The observation of the anionic trithiocarbonate furthermore confirms steps (ii)–(iv) outlined in Fig. 3. Here we also proposed that c5c could be generated through backbiting reaction following BO insertion of the trithiocarbonate intermediate and this explains its formation in this model experiment. Unfortunately, MTCK is only sparingly soluble in organic solvents which prevents reactivity studies. Combined our results show that LiOBn serves the role of a catalyst rather than a mere initiator and that the metal choice is crucial for controlling the O/S exchange process which likely occurs on a kinetic basis. Nevertheless, many questions remain unanswered, and a more detailed mechanistic study is currently underway.Open in a separate windowFig. 6(Left) Selected regions of the 1H NMR spectra (CDCl3, 25 °C, 500 MHz) of polymers corresponding to the polymers obtained from ). The obtained materials are amorphous in nature (Tg = 33.7 °C, for 64–66 The polymers show excellent solubility in organic solvents (THF, DCM, CHCl3) even at high molecular mass, an attractive processing property, and the high molecular mass materials (20 The terpolymers can be formally considered polyesters with regularly distributed trithiocarbonate links and this should result in properties typical for sulfur containing polymers such as susceptibility to oxidation and photolysis.25,36,67,68 Indeed, we find that irradiation of the ROTERP polymer with broadband UV light or dispersion in H2O2 leads to selective cleavage of the trithiocarbonate groups (99% cleavage for 16 h UV irradiation or 5 d dispersion in aqueous H2O2) and degrades the material into oligomers with Mn < 1 kDa (ESI Section S7). The 1H NMR spectra of the product mixtures after degradation show the complete disappearance of the CH2R2-trithiocarbonate groups at ca. 3.75 ppm while CHR3-ester groups at ca. 5.5 ppm can still be detected. To further investigate whether degradability stems from the trithiocarbonate links we prepared a related polyester (without interspersed trithiocarbonate) poly(propylene-orthoterphthalate) via phthalic anhydride/PO ROCOP, and this polymer shows no appreciable degradation under the same conditions. This might imply that there are some degradability benefits to ROTERP polymers over more conventional polyesters as photolysis and oxidation represent the initial breakdown pathways of polymer waste in nature.3 The ROTERP polymer shows furthermore a good refractive index of 1.62 which is similar to that of the parent ROCOP polymers (1.60 for PTA/PO ROCOP polymer and 1.70 for CS2/PO polymer) and this is also typical for sulphur containing polymers.37Having established that ROTERP is a useful methodology for material synthesis we were intrigued to see whether it also allows the synthesis of more complex block polymers structures. In ROCOP, the concept of switchable catalysis has been established as a mechanistically elegant and practical tool to synthesise block-polymers with useful material properties.69–73 Here a suitable catalyst first mediates the ROP of for example cyclic esters with epoxides present in the mixture until the second ROCOP monomer (e.g. CO2) is added causing immediate termination of ROP and the onset of (e.g. CO2/epoxide) ROCOP to form a ROCOP block connected to the ROP polymer.74 As ROTERP formally derives from ROCOP we hypothesised that switchable catalysis between ROP and ROTERP might be possible.To investigate this concept, we first had to identify a suitable ROP that is also mediated by LiOBn in epoxide solvent to then proceed to mechanistic switching. We found that ε-decalactone (εDL) smoothly undergoes living LiOBn catalysed ROP in BO as the solvent without any epoxide ring-opening occurring alongside εDL ROP (ESI Fig. S45–S48). The polymerisation follows a first order rate law with respect to εDL consumption and an excellent initial TOF of 490 h−1 (at 1 eq. LiOBn : 100 eq. εDL : 500 eq. BO and 25 °C) yielding narrow (Đ < 1.2) poly(decalactone) (PDL). Addition of CS2 (500 eq. per LiOBn) and PTA (70 eq.) to polymerising εDL (50 eq) in BO (250 eq.) after 15 min at room temperature completely and immediately stops the occurrence of εDL ROP (Fig. 7). Heating to 80 °C initiates ROTERP and a poly(ester-alt-ester-alt-trithiocarbonate) block grows from the PDL-chain-end until the reaction is stopped after 30 min. Following the polymerisation progress by 1H NMR over time shows the formation of OBn initiated PDL which is followed by a ROTERP block forming uniformly in the second phase of the polymerisation as for the stand-alone ROTERP reactions discussed above. Under these conditions ROTERP occurs in 98% linkage selectivity with 2% erroneous thioester links.Open in a separate windowFig. 7εDL ROP to ROTERP switchable catalysis sequence and 1H NMR spectra (CDCl3, 25 °C, 400 MHz) of aliquots removed at different stages of switchable catalysis. X = OBn.Switchable catalysis and block polymer formation were established by a combination of analytical methods: (i) no εDL is consumed during ROTERP (Fig. 7) and the 13C{1H} PDL C Created by potrace 1.16, written by Peter Selinger 2001-2019 O resonance at δ = 173.2 ppm remains unaffected by the ROTERP process (ESI Fig. S53) showing the cessation of ROP and the absence of transesterification processes between blocks; (ii) the number averaged molecular mass shifts from Mn = 8.40 to 22.69 kDa (Fig. 8), which shows the growth of existing chains rather than the initiation of new ones. This increase in Mn furthermore fulfils statistical considerations for blockpolymer formation;75 (iii) 31P end group analysis shows the consumption of all PDL end groups (ESI Fig. S55);76 (iv) the composition of the resulting block-polymer remains unchanged through multiple precipitations from DCM/MeOH and THF/pentane supporting that the blocks are joint; (vi) DSC analysis shows two Tg''s at −45.7 °C for the ROP block and 26.2 °C for the ROTERP block suggesting microphase separation in the solid state;77 (vii) TGA analysis shows a stepwise thermal decomposition profile with two Td,onset at approximately 205.0 °C for the ROTERP block and 300 °C for the ROP block (both Fig. 8). Previously reported switchable catalyses are associated with a change in the catalytic resting state as shown via in situ UV-VIS and NMR.47,55,78 This is a prerequisite as any active alkoxide chain ends present during ROCOP would lead to the occurrence of ROP. Similarly, we find for our new switches starting from ROP that the addition of the ROTERP monomers causes the immediate emergence of VIS bands at ca. 440 nm prior to any polymer formation (visible as a yellow discolouration of the previously colourless mixture, ESI Fig. S56 and S58). This band is diagnostic for the (C Created by potrace 1.16, written by Peter Selinger 2001-2019 S) chromophore and likely due to the formation of thiocarboxylates. Another indicator for a transformation of the chain-ends is observed in the in situ7Li NMR spectra which is sharpening and shifting by 0.2 ppm upon comonomer addition to polymerising εDL in BO (ESI Fig. S57) and both findings substantiate a change of the catalytic resting state. Together our experiments suggest successful switchable catalysis and block-polymer formation via mechanistic switching from ROP to ROTERP.Open in a separate windowFig. 8Overlayed SEC traces (top left) before and after switch as well as TGA (top right), DSC (bottom left) and 1H–13C HMBC (CDCl3, 25 °C, bottom right) of the obtained block-polymer.In conclusion, we have expanded the repertoire of heteroatom containing polymerisation methodologies by sequence selective ring-opening terpolymerisation (ROTERP). Here three monomers, propylene/butylene oxide A, phthalic thioanhydride B, and CS2 C are enchained by a simple lithium catalyst in an (ABA′C)n fashion. We obtained poly(ester-alt-ester-alt-trithiocarbonate)s with molecular masses of up to >105 g mol−1 that are not easily accessible through other polymerisation methodologies. This unusual insertion selectivity is enabled by a central O/S exchange reaction at the polymer chain-end and we could confirm this hypothesis in model reactions. Lithium is key to achieving high selectivity and activity due to its'' oxophilicity and rigid coordination sphere. Mechanistic experiments also indicate that ROTERP is a kinetically controlled process. With respect to the material properties, we found that incorporation of trithiocarbonate links renders these polymers oxidatively and photodegradable, while showing enhanced thermal stability and solubility compared to some of the related ROCOP polymers. Finally, we demonstrated, that ROTERP is mechanistically compatible with εDL ROP enabling mechanistic switching from ROP to ROTERP for blockpolymer synthesis. We believe that ROTERP is a generalizable methodology with many more viable monomer combinations to be discovered that lead to sequence selective rather than statistical terpolymerisation. ROTERP bears further promise as it can be more selective and faster than the respective ROCOPs. The methodology is mechanistically compatible with ROP and hence can be used for blockpolymer synthesis yielding chemically complex polymer architecture with tuneable material properties. Such materials are now more in demand than ever before given the sustainability challenges our current polymer economy is facing.  相似文献   

20.
We report here a sequential enantioselective reduction/C–H functionalization to install contiguous stereogenic carbon centers of benzocyclobutenols and cyclobutanols. This strategy features a practical enantioselective reduction of a ketone and a diastereospecific iridium-catalyzed C–H silylation. Further transformations have been explored, including controllable regioselective ring-opening reactions. In addition, this strategy has been utilized for the synthesis of three natural products, phyllostoxin (proposed structure), grandisol and fragranol.

We report here a sequential enantioselective reduction/C–H functionalization to install contiguous stereogenic carbon centers of benzocyclobutenols and cyclobutanols.

Molecules with inherent ring strain have gained considerable interest in the synthetic community.1 Among them, four-membered ring molecules have been recognized as powerful building blocks in organic synthesis.2 Driven by ring strain releasing, the reactions of carbon–carbon bond cleavage have been extensively studied in recent years.3 Meanwhile, cyclobutane motifs represent important structural units in natural product and bioactive molecules as well (Scheme 1).4 Therefore, a general and robust method to constitute four-membered ring derivatives is of great value, especially in an enantiomerically pure form.5Open in a separate windowScheme 1Representative cyclobutane-containing bioactive molecules.[2 + 2]-Cycloaddition6 and the skeleton rearrangement reaction7 are two primary methods to prepare chiral cyclobutane derivatives. Recently, the precision modification of four-membered ring skeletons to access enantioenriched cyclobutane derivatives has attracted emerging attention. Several strategies have been developed, including allylic alkylation,8 α-functionalization,9 conjugate addition10 and C–H functionalization11 of prochiral or racemic cyclobutane derivatives (Scheme 2a).12 However, the enantioselective synthesis of chiral benzocyclobutene derivatives is still underdeveloped.13 Although two efficient palladium-catalyzed C–H activation strategies have been developed by Baudoin14 and Martin15 groups via similar intermediate five-membered palladacycles, no enantioenriched benzocyclobutene derivative has been prepared by employing the above two methods. In 2017, Kawabata reported an elegant example of asymmetric intermolecular α-arylation of enantioenriched amino acid derivatives to afford benzocyclobutenones with tetrasubstituted carbon via memory of chirality (Scheme 2b).16 In 2018, Zhang reported an iridium-catalyzed asymmetric hydrogenation of α-alkylidene benzocyclobutenones in good enantioselectivities (3 examples, 83–88% ee).12c To the best of our knowledge, there is no report on enantioselective synthesis of benzocyclobutene derivatives with all-carbon quaternary centers.Open in a separate windowScheme 2Asymmetric synthesis of cyclobutanes and their derivatives. (a) Enantioselective functionalization of four-membered ring substrates. (b) Synthesis of chiral benzocyclobutenone via memory of chirality. (c) This work: sequential enantioselective reduction/C–H functionalization.In line with our continued interest in precision modification of four-membered ring skeletons,9d,10c,12a we initiated our studies on the synthesis of chiral benzocyclobutenes via enantioselective functionalization of highly strained benzocyclobutenones. It is well known that benzocyclobutene derivatives are labile to undergo a ring-opening reaction to release their inherent ring strains.17 Therefore, it is a challenging task to modify the benzocyclobutenone and preserve the four-membered ring skeleton at the same time. We envisioned that a carbonyl group directed C–H functionalization18 of the gem-dimethyl group could furnish enantioenriched α-quaternary benzocyclobutenones (Scheme 2c). This could be viewed as an alternative approach to achieve the alkylation of benzocyclobutenone, which was otherwise directly inaccessible using enolate chemistry through the unstable anti-aromatic intermediate.19 In addition, a highly regioselective C–H activation would be required to functionalize the methyl group instead of the aryl ring. Here we report our work on sequential enantioselective reduction and intramolecular C–H silylation to provide enantioenriched benzocyclobutenols and cyclobutanols with all-carbon quaternary centers. The excellent diastereoselectivity and regioselectivity of silylation were attributed to rigid structural organization of the 4/5 fused ring. Furthermore, this strategy has been utilized to accomplish the total synthesis of natural products phyllostoxin (proposed structure), grandisol and fragranol.We commenced our studies with enantioselective reduction of readily prepared dimethylbenzocyclobutenone 1a (Scheme 3).15,20 Surprisingly, enantioselective reduction of the carbonyl group of cyclobutanone derivatives received little attention. The first reduction of parent benzocyclobutenone was studied in 1996 by Kündig using chlorodiisopinocamphenylborane21 or chiral oxazaborolidines (CBS reduction),22 and only moderate enantioselectivity (44–68% ee) was obtained.23 Although copper-catalyzed asymmetric hydrosilylation of benzocyclobutenone 1a using CuCl/(R)-BINAP gave the benzocyclobutenol ent-2a in 88% ee, optimization of ligands gave no further improvement (Scheme 3a, see Tables S1–S4 for details).24 Gladly, excellent enantioselective reduction could be achieved in 94% yield and 97% ee under Noyori''s asymmetric transfer hydrogenation conditions (Scheme 3b, conditions A, RuCl[(S,S)-Tsdpen](p-cymene)).25 The product 2a showed remarkable stability and no ring-opening byproduct 2a′ was observed. The reduction of parent benzocyclobutenone was examined under conditions A, and benzocyclobutenol was obtained in 90% yield and 81% ee. Apparently, the steric influence imposed by the α-dimethyl group enhanced the enantioselectivity of the reduction. Similarly, the CBS reduction ((S)-B–Me) of benzocyclobutenone 1a gave better results compared with parent benzocyclobutenone, affording the product 2a in 86% yield and 92% ee (Scheme 3c).Open in a separate windowScheme 3Enantioselective reduction of benzocyclobutenone 1a. (a) Copper hydride reduction. (b) Ru-catalyzed asymmetric transfer hydrogenation. (c) CBS reduction.We then examined the substrate scope of the reduction reaction (26 was chosen to improve the yield and enantioselectivity. Besides, benzocyclobutenol 2g with nitro substitution could be obtained in 96% yield and 93% ee. Treatment of pyrrolidinyl substituted benzocycobutenone 1h with catalyst (S,S)-Ts-DENEB afforded desired product 2h in 49% yield and 89% ee, together with ring-opening product 2h′ (18%).Enantioselective reduction of benzocyclobutenonesa
Open in a separate windowaConditions A: 1a (0.5–2.0 mmol), RuCl[(S,S)-Tsdpen](p-cymene) (1–2 mol%), HCOOH/Et3N (5/2), rt. All results are corrected to the (S)-catalyst. The ee values were determined by HPLC analysis; see the ESI for more details.b(S,S)-Ts-DENEB (1–2 mol%) was used, rt or 60 °C.3,3-Disubstituted cyclobutanones were also explored (l-selectride gave cis-4i as a single product in 99% yield and 96% ee. The reaction of 3j gave similar results, and enantioenriched cyclobutanols cis-4j could be furnished in 78% yield and 97% ee from ent-trans-4j (98% ee) following the above oxidation–reduction procedure. The absolute configurations of 2a, ent-2j and trans-4i were unambiguously determined by single-crystal X-ray diffraction analysis of their corresponding nitrobenzoate derivative.27Enantioselective reduction of cyclobutanones 3a
Open in a separate windowaConditions B: 3a (1.0–5.0 mmol), (S)-B–Me (10 mol%), BH3·Me2S (0.6 equiv.), THF, rt.b(S)-B–Me (20 mol%), BH3·Me2S (1.0 equiv.).c(−)-Ipc2BCl (1.2 equiv.), THF, −20 °C. (−)-Ipc2BCl = (−)-diisopinocampheylchloroborane.Inspired by powerful and reliable directed C–H silylation chemistry pioneered by Hartwig,28 we envisioned that the transition-metal catalyzed intramolecular C–H silylations of the above alcohols would provide a single diastereomer owing to rigid structural organization. The challenges here are the control of regioselectivity in the cyclization step and inhibition of the ring-opening pathway. Benzocyclobutenol 2a was chosen as a model substrate to study this intramolecular C–H silylation. The transition-metal catalyst system and alkene acceptors were screened (Scheme 4, see Tables S5–S9 for details). Acceptor norbornene (nbe) derivative A gave the optimal yield in the cyclization step (63% NMR yield), and other phenanthroline ligands gave inferior results. The reaction showed remarkable regio- and diastereoselectivity; no silylation of the arene was detected.With optimal intramolecular silylation conditions in hand, sequential hydroxysilylation/C–H silylation/phenyllithium addition reaction of 2a provided desired product 5a in 56% overall yield without any obvious erosion of enantiomeric purity (
Open in a separate windowaConditions C: i. 2a (0.5 mmol), [Ir(COD)OMe]2 (0.05 mol%), Et2SiH2 (1.2 equiv.), THF, 30 °C; ii. [Ir(COD)Cl]2 (2.5 mol%), Me4Phen (6 mol%), A (1.0 equiv.), THF, 100 °C; iii. PhLi, THF, −78 °C; see the ESI for more details.biii. KHCO3 (2.5 equiv.), H2O2 (10 equiv.), THF/MeOH (1 : 1), 50 °C.Open in a separate windowScheme 4Optimization of intramolecular C–H silylation of benzocyclobutenol 1a.Cyclobutanols were examined under optimal conditions as well (27 The diols cis-6h′ and trans-6h′ could be achieved upon treatment of cyclization products with H2O2 instead of phenyllithium. In addition, bicyclic substrates 4i, 4j smoothly furnished the corresponding enantioenriched products cis-6i, 6j and trans-6i, 6j with four contiguous carbon centers in good yields.Stereospecific C–H functionalization of cyclobutanols 4a
Open in a separate windowaReaction conditions: 4 (0.5 mmol), Ru(PPh3)3Cl (0.2 mol%), Et2SiH2 (1.5 equiv.), THF, 35 °C; ii. [Ir(COD)Cl]2 (2.5 mol%), Me4Phen (6 mol%), A (1.0 equiv.), THF, 100 °C; iii. PhLi, THF, −78 °C; see the ESI for more details.bii. [Ir(COD)Cl]2 (5 mol%), Me4Phen (12 mol%).ciii. KHCO3 (2.5 equiv.), KF (2.5 equiv.), H2O2 (10 equiv.), THF/MeOH (1 : 1), 50 °C.dent-cis-4i (70% ee) was used.eent-trans-4i (97% ee) was used.At this point, we conducted further transformations to explore the utilities of the chiral benzocyclobutene derivatives (Scheme 5). The oxidation of benzocyclobutenol 5a afforded benzocyclobutenone 7 smoothly using Dess–Martin periodinane. This product could be viewed as the result of the alkylation of α-substituted benzocyclobutenone via elusive enolate intermediate I.Open in a separate windowScheme 5Further transformations of benzocyclobutenol 5a.Subsequent Tamao–Fleming oxidation29 with a concomitant cyclobutanone oxidation provided alcohol 8 in 57% yield, albeit with partial loss of enantiopurity. Furthermore, the regioselective Bayer–Villiger oxidation of 7 was achieved using MMPP,30 giving phthalide 9 in 63% yield and 97% ee. Poor regioselectivity was observed when parent benzocyclobutenone was treated with a base.31 In contrast, exposure of 7 to sodium methoxide afforded phenylacetic acid derivative 10 as a single product in 94% yield and 97% ee via proximal bond cleavage.Phyllostoxin (11) was isolated from fungal pathogen Phyllosticta cirsii, and it could represent a potential natural herbicide (Scheme 6).32 The structure was proposed to contain chiral α-quaternary benzocyclobutenone moiety. We envisioned that our strategy would provide a straightforward way to assemble the quaternary center of benzocyclobutenone, thereby confirming the proposed structure and determining the absolute configuration. Our synthesis commenced with enantioselective transfer hydrogenation of substrate 1o. Enantioenriched benzocyclobutenol 2o could be obtained in 93% yield and 99% ee using catalyst (R,R)-Ts-DENEB. Standard procedure, including hydrosilylation/C–H silylation/oxidation, provided diol 5o′ in 89% overall yield and 99% ee. Various oxidation conditions were examined to oxidize diol 5o′, including Swern oxidation, Dess–Martin periodinane and PCC; unfortunately, the reaction only gave messy mixtures. Thus we turned to selective protection of the diol. Selective benzoylation could be achieved via three-step manipulation, giving primary alcohol 12 in 82% overall yield. Swern oxidation and nucleophilic addition of EtMgBr, followed by global deprotection, provided triol 13 in 54% yield over 3 steps. Of mention, benzoyl migration was observed in the EtMgBr addition step. Finally, selective acylation of the phenol and subsequent oxidation furnished benzocyclobutenone 11 in 39% overall yield. However, the optical rotation and NMR spectral data did not match those reported for the natural product.Open in a separate windowScheme 6Total synthesis of the proposed structure of phyllostoxin. Conditions: [Ir(COD)OMe]2, Et2SiH2, THF, rt; ii. [Ir(COD)Cl]2, Me4Phen, A, THF, 100 °C; iii. KHCO3, H2O2, THF/MeOH (1 : 1), 50 °C.The monoterpene grandisol (14) was known as a main component of the sex pheromone of the cotton boll weevil, Anthonomous grandis Boheman, and other insects.33,34 The diastereomer fragranol (15) was isolated in many essential oil aerial parts of plant species such as Achillea fragrantissima, A. falcata and Geranium tuberosum.33 Surprisingly, in comparison to grandisol, there is only one report on enantioselective synthesis of fragranol yet.35 We postulated that our strategy would enable a divergent synthesis of these two diastereomers, starting from an optical resolution of cyclobutanone 3k (Scheme 7). As expected, the CBS reduction of 1x provided cyclobutanols cis-4k and trans-4k (90% yield, 1 : 1.1 dr, 90–99% ee). Subsequent C–H functionalization and oxidation gave diastereomers cis-6k′ and trans-6k′ in good yield. And both diastereomers could be easily separated by column chromatography. Debenzylation, selective silylation of the primary alcohol and Barton–McCombie deoxygenation provided cyclobutanes 17 and 20 uneventfully. Starting from cyclobutane 17, deprotection and subsequent oxidation afforded lactone 18 in 56% overall yield, which led to formal total synthesis of (−)-grandisol 14. Starting from cyclobutane 20, regioselective dehydration with Martin sulfurane and removal of the TBS group furnished alkene 21 in 70% overall yield. Finally, (−)-fragranol 15 was obtained in three additional steps, which included oxidation to an aldehyde, olefination/hydrolysis and reduction.Open in a separate windowScheme 7Divergent synthesis of grandisol and fragranol. Conditions: Ru(PPh3)3Cl, Et2SiH2, THF, 35 °C; ii. [Ir(COD)Cl]2, Me4Phen, A, THF, 100 °C; iii. KHCO3, H2O2, THF/MeOH (1 : 1), 50 °C.  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