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 65 毫秒
1.
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   

2.
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   

3.
Heterogeneous Fenton-like processes are very promising methods of treating organic pollutants through the generation of reactive oxygen containing radicals. Herein, we report novel 0D–1D hybrid nanoarchitectonics (necklace-like structures) consisting of FeCo@N–C yolk–shell nanoreactors as advanced catalysts for Fenton-like reactions. Each FeCo@N–C unit possesses a yolk–shell structure like a nanoreactor, which can accelerate the diffusion of reactive oxygen species and guard the active sites of FeCo. Furthermore, all the nanoreactors are threaded along carbon fibers, providing a highway for electron transport. FeCo@N–C nano-necklaces thereby exhibit excellent performance for pollutant removal via activation of peroxymonosulfate, achieving 100% bisphenol A (k = 0.8308 min−1) degradation in 10 min with good cycling stability. The experiments and density-functional theory calculations reveal that FeCo dual sites are beneficial for activation of O–O, which is crucial for enhancing Fenton-like processes.

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

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

4.
5.
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   

6.
Following immense growth and maturity of the field in the past decade, native mass spectrometry has garnered widespread adoption for the structural characterization of macromolecular complexes. Routine analysis of biotherapeutics by this technique has become commonplace to assist in the development and quality control of immunoglobulin antibodies. Concurrently, 193 nm ultraviolet photodissociation (UVPD) has been developed as a structurally sensitive ion activation technique capable of interrogating protein conformational changes. Here, UVPD was applied to probe the paratopes of nanobodies, a class of single-domain antibodies with an expansive set of applications spanning affinity reagents, molecular imaging, and biotherapeutics. Comparing UVPD sequence fragments for the free nanobodies versus nanobody·antigen complexes empowered assignment of nanobody paratopes and intermolecular salt-bridges, elevating the capabilities of UVPD as a new strategy for characterization of nanobodies.

Ultraviolet photodissociation mass spectrometry is used to probe the paratopes of nanobodies, a class of single-domain antibodies, and to determine intersubunit salt-bridges and explore the nanobody·antigen interfaces.

Antibodies must possess high specificity toward target antigens to enable recognition and activation of an immune response. Because of this specificity, antibodies or antibody fragments are being increasingly explored for deployment in diagnostic assays, vaccine design and other therapeutic applications.1 As such, unravelling the collection of noncovalent interactions and structural features that endow antibodies with specificity is a primary goal in biomedical research and critical for development of new biotherapeutics.2 Methods to decipher antibody–antigen interactions have advanced significantly in recent years, and well-established methods including X-ray crystallography, mutagenesis techniques, cryo-EM, and hydrogen–deuterium exchange mass spectrometry remain the most versatile methods.2–5 Another new strategy, native mass spectrometry (MS) has emerged as a powerful tool for structural biology, including analysis of macromolecular complexes and antibodies.6–10 In this technique, rapid but gentle ionization and transfer of protein complexes into the gas-phase via electrospray ionization (ESI) of aqueous solutions of near-physiological ionic strength preserves noncovalent interactions, allowing elucidation of ligand binding and stoichiometry of complexes.6–9 Coupled to advanced methodologies including ion mobility (IMS) or tandem mass spectrometry (MS/MS), native MS has proven an innovative strategy for interrogating protein structure, revealing topology, distinguishing conformations, identifying ligand binding sites, and determining folding thermodynamics.6–9,11,12 As one example of a premier application in biotherapeutic development, native MS has been utilized to rapidly measure stoichiometry, heterogeneity, and stability of antibody·antigen complexes.10,13–17Detailed structural analyses by native MS often rely on controlled dissociation or disassembly of protein complexes via ion activation in MS/MS workflows.7,8,18–21 Collision-based dissociation methods comprise the most ubiquitous ion activation techniques and are capable of dismantling protein assemblies into subunits, detaching ligands, and facilitating determination of stoichiometries.7,8,18 However, the slow-heating process of collision induced dissociation (CID) causes protein unfolding and less effective fragmentation of the peptide backbone, limiting characterization of native protein complexes.7,8,18 Significantly more structural information can be acquired by employing alternative ion activation techniques.7,8,18 For example, surface induced dissociation (SID) is a collision-based method that promotes disassembly of protein complexes into subunits through a single high energy collision that minimizes protein unfolding and enables robust characterization of native protein assemblies.20,21 SID has proven especially versatile for the analysis of quaternary structure, as protein complexes disassemble to produce subcomplexes and subunits that reflect the native architecture.20,21 Most impressive, the outcome of SID correlates with the magnitude of protein interfacial area (cleavage of the weakest protein:protein interfaces) and has been integrated into computational modelling workflows to enhance the accuracy of protein assembly predictions, emphasizing the value of native MS/MS for structural interrogation.22–24 Alternatively, electron-based dissociation methods result in cleavage of the protein backbone to produce sequence fragments that are enhanced at surface exposed or flexible protein regions, informing topology (through the preservation of noncovalent interactions that prevent separation of reaction products) and degree of protein disorder.7,8,19,25Also a method sensitive to protein structure, ultraviolet photodissociation (UVPD) is a photon-based ion activation method that has demonstrated exceptional use for structural biology. In particular, propensities of polypeptide backbone cleavages induced by 193 nm UVPD correlate with backbone flexibility and arrangement of non-covalent interactions, an outcome related to the likelihood of separation (and detection) of nascent product ions after a backbone cleavage event. Product ions enmeshed by stabilizing non-covalent interactions are less likely to separate (i.e. UVPnoD), causing an apparent suppression of backbone fragmentation. This correlation empowers characterization of conformational changes induced by point mutations, ligand binding, and protein complexation by UVPD.26–33 Additionally, the high sequence coverage and rapid timescale of photodissociation enable detailed analyses of protein gas-phase structure, even informing proton sequestration with single residue resolution.34 As native MS and UVPD increasingly gain broader utility for new protein applications, development and establishment of strategies to routinely study protein structure become imperative to cement these methodologies as cornerstones in the fields of structural biology and biotechnology that encompass development of new therapeutics, imaging agents, diagnostics and drug delivery agents.One fascinating new class of biotherapeutics are nanobodies, single domain antibodies derived from the variable domain of functional heavy chain antibodies found in camelids and certain shark species.35,36 In contrast to conventional ∼150 kDa heterotetrameric immunoglobulin (IgG) antibodies, nanobodies feature a single peptide chain and overall reduced size of ∼15 kDa. Nanobodies offer high stability, solubility, affinity, and specificity, features that have propelled these single domain antibodies as a valuable alternative biotechnology.37–42 Concurrently, these same features facilitate native MS analysis, for which decreased size, increased solubility and antigen affinity are favorable for rapid and routine analysis of nanobody complexes, circumventing analytical challenges and tedious sample preparations, such as proteolysis and deglycosylation steps,14,17,43 often required for native MS of typical IgG·antigen complexes. One recent native MS study mapped the location of the epitope of influenza A hemagglutinin (HA1) bound to an antibody based on a decrease in backbone cleavages of HA1 when bound in the Ab·2HA1 complex relative to the free HA1 antigen during UVPD-MS analysis.31 This innovative approach motivated our interest in exploring an intriguing inverse strategy to map paratopes of nanobodies. In the present work, native MS and 193 nm UVPD are showcased as a valuable combination for determination of intersubunit salt-bridges and nanobody·antigen interfaces, ultimately localizing nanobody paratopes.Three nanobodies44–46 with distinct proteinaceous antigens of green fluorescent protein (GFP), ribonuclease A (RNAseA), and porcine pancreatic amylase (PPA), referred to here as Gnb,44 Rnb,45 and Anb,46 respectively, were targeted to evaluate native MS and UVPD for characterization of nanobody·antigen complexes. Each of these nanobodies interact with the antigen via differing contributions of the complementarity-determining regions (CDR) 1–3 and framework residues to the protein interface, which vary in surface area from 554 Å2 to 683 Å2 to 1062 Å2 for Rnb, Gnb, and Anb, respectively. These three pairs of nanobodies and respective antigens were first ionized individually using native conditions (Fig. S1), and as nanobody·antigen complexes (Fig. 1). In each case, native MS produced the expected 1 : 1 complex in the full mass spectrum (MS1), in accordance with known crystal structures, in a range of charge states. A single charge state of the antigen-bound (bound state) nanobody was isolated and subjected to 193 nm UVPD (Fig. 1B, D and F), resulting in disassembly to release the free nanobody and antigen as well as sequence fragments from backbone cleavage of each protein, the latter of which will be the focus of UVPD analysis. UVPD mass spectra of each free nanobody are shown in Fig. S2. The corresponding backbone cleavage maps for the free nanobodies and corresponding nanobody·antigen complexes are shown in Fig. S3.Open in a separate windowFig. 1Native MS and UVPD of nanobody·antigen complexes. (A) MS1 of Gnb·GFP complex and (B) UVPD of the 13+ charge state. (C) MS1 of Rnb·RNAseA complex and (D) UVPD of the 11+ charge state. (E) MS1 of Anb·PPA complex and (F) UVPD of the 15+ charge state. Insets display expanded views of m/z regions populated by sequence fragments of low relative abundance. For UVPD, 1 laser pulse at 3 mJ was applied.Backbone cleavages induced by UVPD have been shown to be favored at protein regions with higher flexibility, typically ones less stabilized by networks of non-covalent interactions which might prevent separation and release of fragment ions.26–31,47 Accordingly, comparing abundances of fragment ions produced by free and bound nanobodies should reveal regions in which backbone fragmentation is suppressed or enhanced, thus uncovering those residues involved in stabilizing interactions with the antigen and effectively localizing the paratope. Ten types of fragment ions (a, a + 1, b, x, x + 1, y, y − 1, y − 2, and z, where +1, −1, and +2 indicate the gain or loss of hydrogen atoms) commonly generated by UVPD were monitored across the three nanobodies in both the free and bound states. Among the collection of ions detected, a- and x-type ions are the most prevalent and dominant for the nanobodies and complexes, thus providing the greatest sequence coverage. Those fragment ions displaying statistically significant differences in abundances (p < 0.05, n = 5) upon complexation of the nanobody to the antigen are shown in Fig. S4–S6 and mapped onto the backbone position cleaved to generate the fragment ions. The high diversity of fragment types characteristic of UVPD originates from competing pathways: direct dissociation from excited electronic states yields a/a + 1/x/x + 1 ions, and internal conversion to the ground state following intramolecular vibrational energy redistribution (IVR) produces b/y fragments.32,33 IVR processes may preferentially sever weak non-covalent interactions, whereas direct dissociation from excited states occurs on a faster time-scale minimizing disruption of non-covalent interactions.32,33 Consequently, the latter dissociation pathways and respective products, a/a + 1/x/x + 1 ions, are best suited to evaluate antigen-induced changes in nanobody topology. Variations in abundances of these four fragment types tended to be greater at the interface and CDRs, and inspection of the color-coded maps in Fig. S4–S6 reveals that in most cases the abundances of these a/x-type ions decreased for the nanobody·antigen complexes relative to the free nanobodies, signifying suppression of backbone fragmentation.To underscore the impact of antigen binding on UVPD, significant changes in backbone fragmentation (ΔUVPD) based on differences in summed abundances of a/a + 1/x/x + 1 fragments for each free nanobody versus nanobody·antigen complex are shown in Fig. 2, along with color maps highlighting residues at the CDRs and protein interfaces. Generally, apparent suppression of backbone fragmentation (e.g., less efficient separation of nascent fragment ions owing to stabilizing non-covalent interactions) is the greatest at or adjacent to the interface residues, an outcome which is especially notable and consistent for patches of residues near CDR3 in Gnb (Fig. 2A), near CDR1 and CDR3 in Rnb (Fig. 2B), and near CDR2 and CDR3 in Anb (Fig. 2C). These patterns in suppression of backbone fragmentation upon antigen binding correlate with structural features of the respective crystal structures: main antigen contacts in Gnb are predominantly present on CDR3;44 only CDR1 and CDR3 participate in antigen binding for Rnb;45 CDR2 and CDR3 primarily mediate antigen binding for Anb.46 Although nanobodies characteristically feature a disulfide bond spanning C23 to a paired cysteine directly N-terminal to the CDR3, cleavage of the disulfide bond via UVPD was sufficient to unlock the nanobody and allow release of sequence ions from other concomitant backbone cleavages within this region, including CDR1 and CDR2. Moreover, any statistically significant enhancement of backbone cleavages induced by antigen binding was sparse and remote from interaction sites. These infrequent increases in UVPD fragmentation for nanobody·antigen complexes relative to the free nanobody possibly indicate disruption of non-covalent interactions and increased flexibility in those limited regions of the nanobody. However, this sporadic enhancement of backbone cleavages upon antigen binding occurs in stark contrast to more prevalent suppression of backbone cleavages spanning larger swaths of neighboring residues in the nanobody·antigen complexes. Overall, the observed suppression of fragmentation at the protein·protein interface serves as a strong basis for the development of native MS-UVPD strategies to discern nanobody paratopes.Open in a separate windowFig. 2Suppression and enhancement of backbone cleavage sites based on abundances of UVPD fragment ions induced for each nanobody by antigen binding for (A) Gnb, (B) Rnb, and (C) Anb. ΔUVPD heat plots display significant differences (p < 0.05, n = 5) for the abundances of a- and x-type fragment ions between the free and bound nanobody. Blue and red indicate suppression and enhancement of fragment abundances, respectively, for the nanobody upon complexation. Positions that display no significant change are shown in grey, while white indicates small significant changes. Color maps highlighting interface residues and CDRs are also included for each nanobody. ΔUVPD values correspond to the fragment abundance per residue for the bound state minus the fragment abundance per residue for the free state. All product ions had a signal-to-noise ratio > 3.Contrary to classical IgGs where paratopes are primarily localized to CDRs,48 nanobody paratopes feature greater diversity in terms of residue identity and position, and also include the involvement of framework residues as well as the potential absence of interactions from certain CDRs, obfuscating assignment of paratopes.48 By leveraging the trends in the reduction of fragmentation observed in Fig. 2 for the three nanobody·antigen complexes relative to the free nanobodies, a strategy for the approximation of surface patches contributing to the paratope was developed. Because the most structurally significant changes of fragmentation related to antigen binding are demarcated by apparent suppression of backbone cleavages for stretches of neighboring residues, the UVPD data was analyzed by averaging abundances of fragment ions originating from backbone cleavages across every five residues (non-overlapping box-car average) prior to comparing fragment abundances for the free and bound states using Welch''s t-test (n = 5). A stringent significance cutoff of p < 0.001 was subsequently applied to limit false assignments of protein sections involved at the interface. Protein sections (5 residues long) displaying significant suppression of backbone cleavages (i.e. reduction in abundances of fragment ions originating from backbone cleavages in each 5 residue segment) according to this method are plotted onto the crystal structures in Fig. 3. Impressively, only protein sections adjacent to the protein·protein interfaces displayed significant suppression for Gnb and Anb. Similarly, Rnb mostly featured suppression adjacent to the interface, while only two sections remote from the interface (spanning residues 15–24 and 45–49) were also suppressed. Furthermore, a significant enhancement of fragmentation was only observed for Rnb near the N-terminus and C-terminus (Fig. S7), remote from the paratope. This Rnb complex featured the smallest interfacial area, which may lead to instability, compaction, or distortion during ion transmission, resulting in the unexpected suppression and enhancement at these regions. Regardless, UVPD suppression was predominant at the interfaces of each of the three nanobody complexes, presenting a new strategy for the approximation of nanobody residue patches that contribute to the paratope.Open in a separate windowFig. 3The a and x fragment ion abundances originating from backbone cleavages were averaged across every 5 residues for the free and bound forms of each nanobody. Sections displaying significant UVPD suppression upon complexation (p < 0.001, n = 5) were mapped onto the crystal structure as green spheres for (A) Gnb, (B) Rnb, and (C) Anb. Residue positions displaying significant suppression of backbone cleavages, interface residues, and CDR regions are shown for each nanobody as color maps. For comparative purposes, the CDR regions of each nanobody are demarcated on the crystal structures in Fig. S8.Tracking charge states of specifically a- and x-type ions produced by 193 nm UVPD also informs proton sequestration along the protein sequence,34 a feature that is capitalized on here to identify the formation of inter-subunit salt-bridges. In this strategy, charge states for each detected a- or x-type ion are weighted based on intensity and plotted for each backbone cleavage position of the nanobody. For example, if both a42+ and a43+ have equal intensities, the weighted average charge state at backbone cleavage position 4 (between residues 4 and 5) would be 2.5. Because electrostatic interactions, such as salt-bridges, are highly stable in the gas-phase,49,50 applying this method to monitor changes in proton sequestration across free and bound states promises to reveal the locations of inter-subunit salt bridges introduced by complexation, if charge partitioning during subunit ejection is due to heterolytic scission of salt-bridges as previously proposed.26,51,52 Indeed, this is demonstrated in the analysis of the Rnb·RNAseA complex, for which one inter-chain salt bridge has been identified by X-ray crystallography between nanobody R107 and antigen E111. Charge site analysis of the a-ion series for free Rnb (6+ charge state) (Fig. 4A) displays discrete step-changes in charge states at residues R39, between residues 43–47 (suggesting protonation at R45), R68, Q111, and H128, indicating localized protons at these sites. Although coverage of the x-ion series is sparser, it nonetheless corroborates proton localization at Q111 as well as near the N-terminus in the span of residues Q2–L5. Protonation of side-chains is not unexpected for basic amino acids like R and K, but protonation of backbone heteroatoms is also possible for non-basic residues like Q at either the amide oxygen or the amide nitrogen of the peptide bond.53,54 The step analysis reported here localizes all 6 protons of the 6+ charge state of free Rnb.Open in a separate windowFig. 4Weighted average charge of a-type and x-type fragment ions attributed to Rnb produced by UVPD of (A) free Rnb (6+ charge state) and (B) Rnb·RNAseA (11+ charge state), delineated based on the backbone cleavage site along the sequence of the nanobody.Charge site analysis of Rnb in the Rnb·RNAseA complex was not as comprehensive, but nonetheless many charge sites were elucidated (Fig. 4B). Specifically, R29, R46, and R68 remained protonated, while a shift occurred from protonation at Q111 in free RNAseA to protonation of the span of residues 108–106 in the Rnb·RNAseA complex, indicating proton sequestration at R107, a residue engaged in a salt-bridge with the antigen in solution. Charge migration observed upon antigen binding based on this charge-site analysis thus evinces formation of salt-bridges between binding partners. Similarly for Gnb·GFP, charge site analysis enabled localization of multiple charges in both the free and bound states (Fig. S9) including protonation at R36 on Gnb, which is involved in electrostatic interactions with E142 on the GFP antigen. Additionally, a charge was located at R58 for bound Gnb that is absent for free Gnb. Although slightly higher than canonical distance cutoff of 4 Å for salt bridges, Gnb R58 is within 4.5–5.5 Å of the E172 and D173 side-chains of GFP according to the crystal structure. Salt-bridge formation spanning this distance may be possible in the gas-phase, according to charge site analysis derived from the UVPD data. The companion residues, E45 and E104, of the nanobody are also engaged in putative interchain salt-bridges when bound to GFP based on the X-ray structure, however, localization of deprotonation sites of acidic residues is not possible in the positive ion mode.For Anb, coverage of a- and x-type fragments from the Anb·PPA complex was less comprehensive and precluded detailed charge site analyses (Fig. S10). Regardless, data for Gnb and Rnb demonstrate that basic residues engaged in inter-subunit salt-bridges maintain a proton upon disassembly of the nanobody·antigen complexes, highlighting an exceptional use of MS/MS to localize gas-phase salt-bridges between protein subunits. However, it is also noted that these changes in protonation sites may be caused by reasons other than those postulated here. Specifically, it is possible that proton migration is caused by vibrational redistribution of deposited energy from photoabsorption, leading to proton mobilization,53,54 and is not due to heterolytic scission of salt bridges. These results are nonetheless intriguing and may guide future interpretations of charge partitioning during UVPD of protein complexes.Following incredible advances in native MS, routine and rapid analysis of biotherapeutics for quality control and drug development has become commonplace.13–17 At the same time, UVPD has been propelled over the past decade as a premier ion activation method capable of uncovering structural features of proteins that are not revealed by other MS/MS methods.47 As shown in this study, we extended the combination of native MS and UVPD to characterize nanobody·antigen complexes, particularly aiming to showcase UVPD for mapping the binding interface. The pattern of fragment ions generated by UVPD for nanobody complexes resulted in the discernment of inter-chain salt-bridges. Additionally, tracking trends in suppression of fragmentation enabled the localization of nanobody paratopes using only micromolar quantities of nanobodies and circumventing some of the limitations of traditional structural biology methods such as NMR and X-ray crystallography. We anticipate that further improvements in data analysis/informatics methods will further extend the level of structural detail gleaned from the very dense UVPD mass spectra generated for large macromolecular assemblies akin to nanobody·antigen complexes.  相似文献   

7.
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   

8.
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   

9.
10.
Alkynes are widely used in chemistry, medicine and materials science. Here we demonstrate a transition-metal and photocatalyst-free inverse Sonogashira coupling reaction between iodoalkynes and (hetero)arenes or alkenes under visible-light irradiation. Mechanistic and computational studies suggest that iodoalkynes can be directly activated by visible light irradiation, and an excited state iodoalkyne acted as an “alkynyl radical synthetic equivalent”, reacting with a series of C(sp2)–H bonds for coupling products. This work should open new windows in radical chemistry and alkynylation method.

A transition-metal and photocatalyst-free, photoinduced inverse Sonogashira coupling reaction was developed. Under visible-light irradiation, the excited state iodoalkyne acted as an “alkynyl radical synthetic equivalent”.

Alkynes are among the most important class of compounds in organic chemistry. Because of their structural rigidity, special electronic properties and numerous methods available for the functionalization of the triple bond, alkynes are important tools and structural elements both in medicinal chemistry and materials sciences.1 Therefore, the development of a new methodology to introduce carbon–carbon triple bonds is of great importance in organic chemistry. The Sonogashira coupling reaction is typically used for the formation of C(sp)–C(sp2) bonds starting from hetero(aryl) halides and terminal alkynes.2 Recently, “inverse Sonogashira coupling” involving the direct alkynylation of unreactive C(sp2)–H bonds with readily available alkynyl halides has received growing interest in the development of a complementary strategy (Fig. 1a). Various main-group and transition metals have been developed to promote this transformation.3 In addition, a photomediated Sonogashira reaction without a photocatalyst was also developed by several groups (Fig. 1b).4Open in a separate windowFig. 1Models of alkynylation. (a) Conventional inverse Sonogashira reaction. (b) Photomediated Sonogashira reaction. (c) SOMOphilic alkynylation. (d) Photoinduced inverse Sonogashira reaction.In recent years, SOMOphilic alkylnylation (SOMO = singly occupied molecular orbital) has become an excellent method of introducing alkynyl groups (Fig. 1c).5 Based on photoredox and transition metal catalysis, numerous in situ generated radicals undergo α-addition and β-elimination to alkynyl reagents, like the broadly applicable ethynylbenziodoxolone (EBX) reagent. Various radical alkynylations were thus discovered by Li,6 Chen,7 Waser,8 and many other groups.9 However, extending the scope of radical precursors, more atom–economic reactions, and a deeper understanding of the mechanism in these transformations are still highly desirable.After the discovering of trityl radicals by Gomberg in 1900, the “rational” era of radical chemistry has since begun.10 Now, the development of radical reactions, especially those involving C(sp3) and C(sp2) radicals, enables rapid access to drug discovery, agrochemistry, materials science, and other disciplines.11 However, the C(sp) radical remains a baffling species. Due to their very high energy, short life time, and limited and harsh preparation methods, alkynyl radicals remain an elusive species, which just exists in some extreme environments, like outer-space and the petrochemical industry.12 Even though alkynyl radicals have been proposed as intermediates for some alkynylation methods, they were regarded as mysterious species and ignored by organic chemists for a long time.13 Recently, two approaches have been developed to aid the alkynyl radical generation step. In 2015, Hashmi and collaborators reported a [Au2(μ-dppm)2]2+ catalyzed free radical–radical C(sp)–C(sp3) bond coupling reaction between iodoalkynes and aliphatic amines.14 Under irradiation of sunlight, the dimeric gold complex was proposed to reduce the iodine acetylide to an alkynyl radical. In 2017, Li developed a transition-metal-free alkynylation reaction between iodoalkyne and 2-indolinone.15 Iodoalkynes could release alkynyl radicals under high temperature conditions. In 2019, we reported an Au(i) and Ir(iii) catalyzed alkynylative cyclization of o-alkylnylphenols with iodoalkynes, wherein the photosensitized energy transfer promoted the oxidative addition of a gold(i) complex with iodoalkynes.16 Based on our continuous interest in haloalkyne and photo-chemistry, we proposed that an iodoalkyne could be a potential “alkynyl radical precursor” under light irradiation. In this work, we uncovered a novel mode of transition-metal and photocatalyst-free, direct photoexcitation of iodoalkynes for the inverse Sonogashira coupling reaction with arenes, heteroarenes, and alkenes via an “alkynyl-radical type” transfer (Fig. 1d).  相似文献   

11.
The OH + HCl → H2O + Cl reaction releases Cl atoms, which can catalyze the ozone destruction reaction in the stratosphere. The measured rate coefficients for the reaction deviate substantially from the Arrhenius limit at low temperatures and become essentially independent of temperature when T < 250 K, apparently due to quantum tunneling; however, the nature of the quantum tunneling is unknown. Here, we report a time-dependent wave packet study of the reactions on two newly constructed potential energy surfaces. It is found that the OH + HCl reaction possesses many Feshbach resonances trapped in a bending/torsion excited vibrational adiabatic potential well in the entrance channel due to hydrogen bond interaction. These resonance states greatly induce quantum tunneling of a hydrogen atom through the reaction barrier, causing the reaction rates to deviate substantially from Arrhenius behavior at low temperature, as observed experimentally.

The OH + HCl reaction possesses many Feshbach resonances trapped in the hydrogen bond well in the entrance channel, which substantially enhance the reaction rates at low temperatures.

In the classical picture, a chemical reaction with an energetic barrier can only occur at collision energies higher than the barrier, which leads to the well-known Arrhenius formula for chemical reaction rates. However, chemical reactions can happen at energies below the reaction barrier through quantum tunneling,1–3 resulting in the deviation of the reaction rates from the Arrhenius behavior at low temperatures. The effect of quantum tunneling on the reaction rates increases with decreasing reaction temperature, and hence becomes especially important in low-temperature environments such as the interstellar medium and atmospheric processes.4,5 Reaction resonances are quasi-trapped quantum states in the transition state region with some lifetime, and can substantially promote quantum tunneling through the reaction barrier. Over the past decades, great efforts have been devoted to detecting resonances in chemical reactions and to studying their structures and dynamics.6–14 Theoretical studies on the O(3P) + HCl → OH + H reaction using the accurate 3A′′ potential energy surface revealed that the tunneling induced by the resonances trapped in the van der Waals well in the reactant channel can substantially enhance the thermal rate constants at low temperatures.15–17 A combined experimental and theoretical investigation discovered that resonance-induced quantum tunneling dramatically enhances the reactivity of the F + p-H2 → HF + H reaction,14 and is fully responsible for the unusually high chemical reactivity of the reaction in the low temperature interstellar medium. Recently, a quantum dynamics calculation showed that the presence of two resonance peaks strongly influence the rotational quenching of HF (j = 1, 2) with H, leading to an up to two-fold increase in the thermal rate coefficients at the low temperatures characteristic of the interstellar medium.18 Thus, understanding quantum tunneling, and in particular, resonance-induced quantum tunneling, in chemical reactions is of general interest and fundamental importance to low-temperature chemistry.The OH + HCl → H2O + Cl reaction is of great importance in atmospheric chemistry because it releases Cl atoms from one of the principal chlorine-containing species in the stratosphere, HCl. The Cl atoms generated from the reaction can catalyze the ozone destruction reaction in the stratosphere, which was responsible for the formation of the ozone hole over Antarctica.19 Since the chlorine-catalyzed ozone destruction is proportional to the steady-state Cl atom concentration, which is directly controlled by the rate of the reaction, extensive studies have been carried out to measure the rate coefficient for the reaction with high accuracy over a wide range of temperatures.20–29 The measured rate coefficients exhibit a small activation energy of a few hundred K, deviate substantially from the Arrhenius limit at low temperatures and become essentially independent of temperature when T < 250 K.25,28,29 In addition, a large H/D kinetic isotope effect has also been found.21,22,26,28 All these observations suggest the presence of an important quantum tunneling effect in the reaction.The dynamics of this reaction and its reverse have also attracted great attention in the past decades. In particular, the endothermic Cl + H2O → HCl + OH reaction with a late barrier has been extensively investigated as a benchmark system for mode specificity and bond selectivity chemistry.30–35 Recently, the construction of two high-quality potential energy surfaces (PESs) in its ground electronic state using the PIP-NN method have substantially advanced the theoretical study of the dynamics and kinetics of the system. The first PES was based on a large number of ab initio data points calculated at the multi-reference configuration interaction (MRCI) level of theory by Li, Dawes, and Guo (LDG),33 and the second one was fitted to ab initio energy points obtained using an explicitly correlated unrestricted coupled-cluster method with single, double, and perturbative triple excitations (UCCSD(T)-F12b) and the augmented correlation-consistent polarized valence triple-zeta (aug-cc-pVTZ, or AVTZ) basis set by Zuo, Zhao, Guo, and Xie (ZZGX) with a fitting error of 6.9 meV.36 The static barrier height is 2.86 and 2.23 kcal mol−1, respectively, for these two PESs. On both PESs, there exists a well of about 3.5 kcal mol−1 in the OH + HCl entrance channel due to the hydrogen bond (HB) interaction between OH and HCl. Extensive quantum dynamics studies on the PESs have revealed many interesting features of the reaction in both directions. In particular, time-dependent wave packet calculations for the title reaction with OH in the ground and vibrational excited states found one or two broad peaks in the total reaction probabilities, which are presumed to be the signature of the resonances supported by the reactant complex well.36,37 However, the impact of these peaks on the reaction rates has not been investigated. Ring-polymer molecular dynamics (RPMD) calculations38,39 were also carried out on both PESs to compute the thermal rate coefficients for the reaction.40–42 It was found that the RPMD rates on the LDG PES underestimate the experimental data, while the RPMD rates on the latest ZZGX PES agree with the experimental results much better, and do not decrease further when the temperature falls below 300 K, apparently due to quantum tunneling. Unfortunately, RPMD calculations cannot provide any clue regarding the nature of the quantum tunneling.Therefore, despite the significant progress that has been made in theoretical studies of the system, some key issues still remain to be addressed: Are the broad peaks in the total reaction probabilities obtained from the time-dependent wave packet calculations indeed the signature of the resonances in the reaction? If not, do there exist resonances in the reaction? How do resonances affect the reaction rates at low temperature? Here, we report a quantum dynamics study of the reaction on two new and more accurate PESs. Good agreement is achieved between the rate coefficients calculated on these two PESs and the experimental data. Our calculations reveal that the HB well in the entrance channel of OH + HCl supports many low energy resonance states. These resonance states substantially enhance the quantum tunneling effect and have an important impact on the reaction rates at low temperatures.In order to improve the fitting accuracy of the ZZGX PES, we constructed two new PESs using the fundamental invariant neural network (FI-NN) method, fitted to ∼70 000 ab initio energy points calculated at the UCCSD(T)-F12a and UCCSD(T)-F12b levels of theory, respectively, both with the AVTZ basis set. The fitting RMSE is 3.07 and 3.12 meV, respectively, for the F12a and F12b PES, which is about half that for the ZZGX PES.36 The spin–orbit coupling of the channels of both the reagent OH and the product Cl have been included using FI-NN fitting to about 38 000 points calculated at the MRCI/aug-cc-pVTZ level of theory with a fitting error of 0.19 meV. The CASSCF wave-function with an active space of (5e, 3o) was used as a reference for MRCI. Details of the new PESs are provided in the ESI. The static barrier height for PESa is 0.088 eV (0.095 eV with SO correction included), and that for PESb is 0.097 eV (0.104 eV). As can be seen from Tables S1 and S2, the geometries and energies of all the stationary points for the F12b PES without SO correction are in good agreement with the ZZGX PES. Table S2 also shows the corresponding complete basis set (CBS) energies for these stationary points based on AVTZ, AVQZ and AV5Z calculations. Because the F12b energies are slightly closer to the CBS results, we will present the dynamical results obtained on the F12b PES in the main text and provide those for the F12a PES in the SI.On the new PESs, we carried out potential-averaged five-dimensional (PA5D) time-dependent wave packet43,44 calculations to obtain the total reaction probabilities for the reaction by freezing the non-reacting OH bond in its ground vibrational state. Tests revealed that the PA5D treatment is capable of providing reaction probabilities for the ground rovibrational initial state that are essentially identical to those obtained using the full six-dimensional approach, as shown in Fig. S2.Fig. 1(A) shows the total reaction probabilities for the HCl + OH reaction as a function of collision energy calculated on the F12b PES with both reagents in the ground rovibrational state at propagation times of 60 000, 120 000, 360 000, and 2 400 000 a.u. At high collision energies, the reaction probabilities converge quickly with respect to the propagation time, and one can barely see any difference among the four reaction probability curves, which exhibit smooth increases with collision energy. However, in the low collision energy region, large differences appear for different propagation times. At t = 60 000 a.u., the reaction probability presents a smooth curve with some small and broad oscillations, as was observed in the wave packet calculations of Guo and coworkers.36Open in a separate windowFig. 1(A) Total reaction probabilities for the ground initial state of the OH + HCl → Cl + H2O reaction on the F12b PES at wave packet propagation times of T = 60 000, 120 000, 360 000, and 2 400 000 a.u. (B) Same as (A), except showing the collision energy between 0.0 and 0.04 eV. The crosses mark the points for which the wavefunctions are shown in Fig. 2. (C) Total reaction probabilities for some partial waves J = 0, 30, 60, and 90 as a function of the collision energy. (D) Same as (C) except showing the collision energy between 0.0 and 0.04 eV.When the propagation time is increased to t = 120 000, the reaction probabilities at collision energies below 0.05 eV increase substantially. With further increasing the propagation time to 360 000 a.u., many oscillatory structures emerge at collision energies below the barrier height of 0.104 eV, in particular in the very low collision energy region as shown in Fig. 1(B). These sharp structures become fully converged after around 2 400 000 a. u. of wave packet propagation (∼58 ps). The reaction probability even at a collision energy close to zero reaches 3–4%. The convergence of the total reaction probabilities on the F12a PES is very similar to that on the F12b PES (Fig. S3), except that the final converged reaction probabilities for these two PESs exhibit a small shift, apparently due to slightly different barrier heights. Therefore, it is clear that reaction resonances exist in the title reaction in the very low collision energy region, and that these resonance states substantially induce quantum tunneling and enhance the reactivity. The lifetimes for these resonance states are quite long, with many being longer than 6.5 ps and having corresponding widths smaller than 0.1 meV. Fig. 1(C) presents converged (t = 2 400 000 a.u. ≈ 58 ps) total reaction probabilities for the total angular momentum J = 0, 30, 60, and 90. With increasing J, the reaction probability curve shifts to higher energy. In the low collision energy region (<0.05 eV), the total reaction probabilities for J = 30 exhibit rich oscillatory structures as in the J = 0 case (Fig. 1(D)), which are expected to have a great influence on the rate constant at low temperature. With further increase of the total angular momentum, the influence of these resonances on the total reaction fades due to the centrifugal barrier, which prevents the low-energy wave function from entering the well. They only leave a small trace in the total reaction probabilities at low energies for J = 60, and do not have any effect for J = 90.To understand the nature of these resonances, we calculated scattering wave functions at two collision energies (1.08 and 4.26 meV) with the peak reaction probabilities indicated by x in Fig. 1(B). Fig. 2(A) shows the two dimensional (2D) contour at the collision energy of 1.08 meV in the Jacobi coordinates HCl bond length (rH–Cl) and center of mass distance between OH and HCl (RHCl–OH), with the other coordinates integrated. As can be seen, the wave function is localized in the HB well region in the entrance channel with a few nodes in the R coordinate and no node in the rH–Cl coordinate. Inspection of the scattering wave function for the bending and torsion coordinates reveals nodes exist in these coordinates (Fig. S4 and S5). The 2D contour in the coordinates of RHCl–OH and rH–Cl at the collision energy of 4.26 meV shown in Fig. 2(B) looks similar to that shown in Fig. 2(A), except with more nodes in the R direction. Therefore, the observed resonance states in the reaction are Feshbach resonances trapped in a bending/torsion excited vibrationally adiabatic potential (VAP) well in the reactant complex region due to the HB interaction.Open in a separate windowFig. 2Reactive scattering wave functions for the OH + HCl → Cl + H2O reaction on the F12b PES in the two Jacobi coordinates R(HCl–OH) and r(H–Cl) with other coordinates integrated at the collision energies of 1.08 (A) and 4.26 meV (B). The contour lines are the corresponding 2D PESs along the two reactive bonds R(HCl–OH) and r(H–Cl) with other coordinates optimized. The geometries for the saddle point and the HB minimum are displayed in (A). The coordinate units in the figures are a0. Fig. 3(A) shows the accurate rate constants for the initial ground rovibrational state, kg, based on the probabilities for J = 0, 30, 60, 90 using the uniform J-shifting approach with a temperature-dependent shifting constant.45–47 A test shows that the J-shifting scheme based on these four individual J values only introduces a few percent error to the rate constants in the temperature region considered here (Fig. S6). As can be seen from the figure, with decreasing temperature, kg first decreases rapidly from 1000 K to 700 K, then decreases slowly. It reaches a minimum at T ≈ 260 K, and increases slowly with further decrease of the temperature. At temperatures lower than 300 K, the rate constants for the ground rovibrational initial state are larger than the measured thermal rate coefficients, with kg being larger than kexp by ∼70% at T = 200 K.Open in a separate windowFig. 3(A) Accurate rate constants, kg, for the initial ground rovibrational state of the HCl + OH → H2O + Cl reaction calculated on the F12b PES, in comparison with kJSg (obtained using the J-shifting approximation) and kNRg (based on the background reaction probabilities for J = 0 with the resonance contribution removed shown below); (B) the background reaction probabilities up to E = 0.1 eV by connecting some valleys of the reaction probabilities marked by x.Also shown in Fig. 3(A) are the rate constants for the ground rovibrational initial state, kJSg, obtained from the J = 0 reaction probabilities using the J-shifting approximation (see ESI for details). As can be seen, the J-shifting approximation works very well at high temperatures around 1000 K, but begins to overestimate the rates with decreasing temperature. At T = 500 K, kJSg is about 10% higher than the true rate. At temperatures below 300 K, the J-shifting approximation underestimates the reaction rate, with kJSg being smaller than kg by about 16% at T = 200 K. Overall, the J-shifting approximation works fairly well for the ground rovibrational initial state, although there are numerous resonance peaks in the reaction probabilities in the low collision energy region.Now we consider the issue of the effect of the resonance structures found in the reaction on the rate constant. For a reaction system with isolated resonances, it is rather straightforward to remove the resonance contributions from the reaction probabilities by fitting the resonance peaks to some Lorentzian functions and to obtain smooth background scattering probabilities, as demonstrated in the F + HD reaction and recently in the inelastic scattering of H + HF (ref. 6 and 18). However, the OH + HCl reaction possesses numerous highly overlapped resonances in the low energy region, as shown in Fig. 1(A), and it is impractical to fit these resonance peaks as accurately as the isolated resonances. Instead, we obtained an approximate background reaction probability curve, which is shown in Fig. 3(B), by smoothly connecting some resonance valleys as shown in the figure. The rate constants kNRg calculated using the background curve shown in Fig. 3(B) with the J-shifting approximation were compared with the original kg in Fig. 3(A). Given the fact that the J-shifting approximation works fairly well for kJSg, as shown in Fig. 3(A), it is very reasonable to expect that will work even better for the kNRg values obtained from the background reaction probabilities with the resonance contribution removed. As can be seen, kNRg exhibits rate behavior typical for systems with a low barrier with some quantum tunneling effects. At T = 1000 K, kNRg is essentially identical to kg, but decreases much faster than kg as the temperature drops. At T = 200 K, kg is larger than kNRg by a factor of about 5.6 (9.8 × 10−13vs. 1.74 × 10−13 cm3 s−1), indicating that the reaction probabilities in the low collision energy region due to resonances substantially enhance the rate constants for the reaction at low temperatures. It is worthwhile to note that for systems with overlapped resonances like that shown in Fig. 3(B), the reaction probabilities at the valleys must be considerably higher than the true background reaction probabilities; therefore, the background curve shown in Fig. 3(B) is the upper limit of the background reaction probabilities and the rate based on the curve shown in Fig. 3(A) is also the upper limit of the rate for the ground rovibrational initial state without the resonance contributions. Therefore, the true enhancement due to the resonances must be larger than that shown in Fig. 3(A).For reliable comparison with the measured thermal rate coefficients, we must take into account the contributions from all the thermally populated initial states of the reagents. Due to the very large number of thermally populated rotational states for this reaction even at 200 K, we opted to calculate the cumulative reaction probabilities, NE(E), (the sum of the reaction probabilities for all the initial states with a fixed total energy) for the reaction from which the thermal rate constants can be reliably evaluated.48–52 The transition state wave packet calculations were carried out to obtain NE(E) using the details given in the ESI. Due to huge computational efforts required to obtain the cumulative reaction probabilities for J > 0, we only calculated NE(E) for J = 0 and employed the JK-shifting approximation45,49,53 to obtain the thermal rate constant. NE(J = 0, E) as a function of total energy measured with respect to the ground rovibrational energy of OH and HCl is presented in Fig. S7.In Fig. 4, we present thermal rate constants for the reaction calculated on both the F12a and F12b PES, together with the rate constants for the ground rovibrational state (kg) and the previous experimental measurements.21,22,24–29 As can be seen, the thermal rate coefficients are smaller than kg over the entire temperature region, indicating that reagent rotation excitations diminish the reaction rates. Overall, the thermal rate coefficients calculated on both PESs agree with the experimental results rather well, with the F12a PES slightly overestimating and the other PES slightly underestimating compared to the experimental measurements. As observed in the experiments, the thermal rate coefficients decrease quickly with decreasing temperature in the high-temperature region, but decrease much more slowly at low temperatures, in particular at T < 300 K, substantially deviating from Arrhenius behavior.Open in a separate windowFig. 4Thermal rate constants of the HCl + OH → H2O + Cl reaction calculated on both the F12a and F12b PES, compared with rate constants for the ground rovibrational state, the RPMD rates on ZZGX PES and the previous experimental measurements.21,22,24–29 Experimental data are taken from Husain et al. (in circles),21,22 Molina et al. (in diamonds),24 Ravishankara et al. (in upward-pointing triangles),25 Smith et al. (in squares),26 Sharkey et al. (in triangle left),27 Battin-Leclerc et al. (in downward-pointing triangles),28 and Bryukov et al. (in leftward-pointing triangles).29Also shown in Fig. 4 are the RPMD rates calculated (see ESI for details) on the F12a, F12b and ZZGX PESs.42 As shown, the RPMD rates on the F12b PES agree with those on the ZZGX PES extremely well, although the barrier heights for the F12b PES are higher by 9 meV. The present RPMD rates agree with the JK-shifting quantum rates rather well, except at 200 K, at which the RPMD rates are overestimated by about 40%. As a result, the RPMD rate is higher than the experimental value at T = 200 K, in particular for the F12a PES. Previous studies have shown that RPMD tends to underestimate reaction rates in the strong quantum tunneling region, even for systems with resonances such as the O(3P) + HCl → OH + H reaction;54 however, it overestimates the rates for the F + H2 reaction with pronounced post-barrier Feshbach resonances55 and some insertion reactions.56 The discrepancy between the RPMD and JK-shifting quantum rate at T = 200 K could be caused by a possible underestimation of the rate by the JK-shifting approximation, as in the J-shifting approximation of the ground rovibrational initial state shown in Fig. 3(A). On the other hand, the rapid increase in the rate with decreasing temperature (from 250 K to 200 K) could also be a feature of bimolecular reaction rates in the high-pressure limit for reactions with a pre-reactive minimum.57 RPMD rate theory in its original form employs a free energy calculation,38,39 which allows thermalization in the pre-reactive minimum, sampling tunneling pathways with energies not accessible in the low-pressure limit. Therefore, RPMD models a rate process that resembles rates in the high-pressure limit, which could lead to an overestimation of the rate at low temperature. A recent study also reported a similar issue with RPMD.58 The authors attributed it to spurious resonances in RPMD favoring energy transfer in the pre-reactive minimum, which is similar to our argument. Certainly, more quantum dynamics calculations to provide rigorous quantum rates at low temperatures to assess the accuracy of the JK-shifting approximation and the RPMD method on this reaction system with strong HB interaction in the entrance channel would be highly desirable.Therefore, the OH + HCl reaction possesses many long-lifetime Feshbach resonances trapped in a bending excited VAP well in the entrance channel due to HB interaction. These resonance states substantially induce quantum tunneling of the hydrogen atom through the reaction barrier and enhance the reactivity in the low-collision region. Consequently, the reaction rate for the reaction becomes essentially independent of temperature in the low-temperature region and deviates substantially from Arrhenius behavior, as observed experimentally. The resonance states in the reaction are very different in location from those in the F + H2/H2O reactions,8,11,12,14,59–63 which are trapped in the VAP well in the product channel. Furthermore, they are trapped in the bending/torsion excited VAP well with HCl in the ground vibrational state, unlike those in the F + H2/H2O reactions with HF in vibrationally excited states.8,63,64 In nature, these resonance states in the reaction arise from HB interaction as in the F + H2O (v = 0) reaction,11,62,63 but are different from those in the F + H2/HD/HOD (v = 1) reactions8,10,12,14,64–66 due to chemical bond softening. Because the resonance states are trapped in the bending/torsion excited VAP well, their lifetimes are much longer than those observed in the F + H2/H2O reactions. As a result, they should have an important impact on the differential cross sections. More efforts, and in particular more joint efforts involving theory and experiment, should be devoted to studying these Feshbach resonances in this reaction of great atmospheric importance in detail.  相似文献   

12.
Directed evolution is a powerful approach to engineer enzymes via iterative creation and screening of variant libraries. However, assay development for high-throughput mutant screening remains challenging, particularly for new catalytic activities. Mass spectrometry (MS) analysis is label-free and well suited for untargeted discovery of new enzyme products but is traditionally limited by slow speed. Here we report an automated workflow for directed evolution of new enzymatic activities via high-throughput library creation and label-free MS screening. For a proof of concept, we chose to engineer a cyclodipeptide synthase (CDPS) that synthesizes diketopiperazine (DKP) compounds with therapeutic potential. In recombinant Escherichia coli, site-saturation mutagenesis (SSM) and error-prone PCR (epPCR) libraries expressing CDPS mutants were automatically created and cultivated on an integrated work cell. Culture supernatants were then robotically processed for matrix-assisted laser desorption/ionization time-of-flight (MALDI-ToF) MS analysis at a rate of 5 s per sample. The resulting mass spectral data were processed via custom computational algorithms, which performed a multivariant analysis of 108 theoretical mass-to-charge (m/z) values of 190 possible DKP molecules within a mass window of 115–373 Da. An F186L CDPS mutant was isolated to produce cyclo(l-Phe–l-Val), which is undetectable in the product profile of the wild-type enzyme. This robotic, label-free MS screening approach may be generally applicable to engineering other enzymes with new activities in high throughput.

A robotic workflow for directed evolution of new enzymatic activities via high-throughput library creation and label-free MS screening.

Mimicking random mutagenesis and natural selection in the laboratory,1–3 directed enzyme evolution is valuable in improving a variety of properties of biocatalysts (i.e., catalytic activity, stability, substrate specificity, and enantioselectivity).4–6 While it is straightforward to generate thousands of enzyme variants using established methods such as error-prone PCR (epPCR), site-saturation mutagenesis (SSM), and DNA shuffling, phenotypic screening remains a major bottleneck due to the substrate and product versatility of biocatalysts.7,8 Optical assays are widely applied in high-throughput screening (HTS) during directed evolution, often requiring a surrogate substrate or a coupling assay with a colorimetric or fluorogenic readout.9,10 However, it is very challenging to apply such approaches to discover new catalytic activities. On the other hand, mass spectrometry (MS) is label-free and suitable for untargeted molecular profiling.11 The high sensitivity and chemical resolution of MS further highlight its potential in discovering new products, because new catalytic activities are often weak when first emerging from enzyme promiscuity.12We and others have recently developed a range of high-throughput MS screening methods for directed protein evolution.13–16 However, the label-free advantage of MS has not been fully demonstrated in engineering new enzymatic activities, possibly due to the difficulties in untargeted MS screening. Particularly, it requires careful standardization and optimization of sample preparation, MS acquisition, and data processing, which are necessary to minimize experimental noise and spot the weak signals of a new product. On the other hand, biofoundries provide an emerging infrastructure to assist the design–build–test–learn (DBTL) cycles in biological engineering via robotic standardization and parallelization.17–19 Using an integrated biofoundry, here we report a workflow for unlabeled MS screening of recombinant libraries to rapidly isolate enzyme mutants that catalyze the formation of new products. This new workflow extends our previous matrix-assisted laser desorption/ionization time-of-flight (MALDI-ToF) MS-based screening approach from agar colonies16 to liquid cultures in industry-standard microplates for better uniformity. Unlike colony biomass, liquid culture media often contain high concentrations of nonvolatile components, which interfere with MALDI matrix crystallization and cause severe ion suppression during MS analysis. Therefore, sample preparation steps, such as liquid–liquid extraction and solid-phase extraction, need to be incorporated and automated.Cyclodipeptide synthases (CDPSs) are a family of enzymes that catalyze the formation of a diketopiperazine (DKP) from two aminoacyl-tRNA substrates (aa-tRNAs), which is an important pharmacophore for modern drug development20–22 (Fig. 1A and S1). To produce new DKP derivatives for therapeutic and industrial applications, the catalytic mechanism of CDPSs has been studied23,24 to guide engineering. AlbC (239 aa) is the first identified CDPS protein, taking phenylalanyl-tRNAPhe (Phe-tRNAPhe) and leucyl-tRNALeu (Leu-tRNALeu) as substrates to synthesize cyclo(l-Phe–l-Leu) (cFL) in its native host Streptomyces noursei.25,26 Recombinant E. coli with AlbC overexpression is able to produce more cyclodipeptide derivatives in addition to cFL.27 To date, eight CDPSs have been structurally characterized,26,28–32 and mutagenesis studies reveal that the residues within the P1 and P2 catalytic pockets are the key determinants of substrate specificity. However, the detailed mechanism of substrate recognition and catalysis is still elusive, and CDPSs are recognized as recalcitrant targets for rational engineering.33,34 To our knowledge, large-scale screening of CDPS variants has not been reported, possibly due to the lack of applicable HTS assays.Open in a separate windowFig. 1(A) Scheme of diketopiperazine (DKP) biosynthesis catalyzed by cyclodipeptide synthase (CDPS). (B) Workflow of the matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-ToF MS)-based high-throughput screening method for directed evolution of AlbC protein in E. coli.Here we develop and apply untargeted MS screening on a biofoundry for directed evolution of AlbC mutants that form new CDP products. The workflow consists of strain library creation, colony picking, microtiter cultivation, organic solvent extraction, transfer of the sample to a MALDI target and acquisition of mass spectra, followed by data processing and visualization (Fig. 1B). For library creation, AlbC variants were generated by either SSM or epPCR approaches and expressed from a pET28a plasmid under control of a T7 promoter. Upon genetic transformation of E. coli Rosetta(DE3), individual clones were transferred by a colony picker into microtiter plates. Subsequent cultivation, inducible production by addition of isopropyl β-d-1-thiogalactopyranoside (IPTG), ethyl acetate extraction, and MS sample preparation were performed using an integrated robotic workcell in the Shenzhen Synthetic Biology Infrastructure. This workcell integrates common instruments for synthetic biology, including a liquid-handling station, shaking incubators, centrifuge, plate reader, and so on. Organic extracts of the liquid cultures were spotted onto a MALDI target and overlaid with 4-CHCA matrix solution. MS targets were then manually transferred to a stand-alone MALDI mass spectrometer in this study, although a robotic configuration has been previously reported to automate this step.35We first applied the workflow to analyze the wild-type (WT) AlbC-expressing strain, Rosetta(DE3)/pET28-AlbC(WT). After cell cultivation, inducible production, and sample preparation, MS analysis was performed using an Autoflex MALDI-ToF mass spectrometer in the reflector positive ion mode (Fig. 1). In the resulting MALDI mass spectra, we observed [M + H]+ ion peaks with m/z values corresponding to cFL (m/z 261.17), cFY (m/z 311.15), cFM (m/z 279.12), cYL (m/z 277.16), cFF(cYM) (m/z 295.15), cLL (m/z 227.16) and cMM (m/z 263.12), all of which were absent from the control strain Rosetta(DE3)/pET28 (Fig. S2). Production of these CDP molecules was further confirmed by examining tandem MS results using LC-MS in the multiple reaction monitoring (MRM) mode (Table S3), where ion fragmentation patterns were consistent with the literature.26,27 Moreover, the observed production profile was also consistent with previous studies.26,27 These results validated our workflow to detect CDP products from microplate cultures of AlbC-expressing E. coli using MALDI-ToF MS.To engineer AlbC mutants that synthesize new products, we first focused on the key residues in the substrate-binding pockets. AlbC employs a “ping-pong” catalytic mechanism: the initial step transfers the aminoacyl moiety of the first aa-tRNA onto a conserved serine, leading to the formation of an aminoacyl enzyme intermediate; then, the aminoacyl enzyme reacts with the aminoacyl moiety of the second aa-tRNA to form a dipeptidyl enzyme.32 Previous biochemical experiments demonstrated that two binding pockets P1 and P2 of AlbC accommodate the aminoacyl moieties of the two aa-tRNA during the biosynthesis24 (Fig. 2). Only a limited number of mutations were examined on P1 and P2 residues,25,26 possibly due to the lack of HTS assays for large-scale analysis. Instead, we applied our workflow to create and screen the SSM libraries of select residues, including 10 residues in the binding pockets (L33, V65, L119, L185, L200, M152, M159, I204, T206 and P207) and 4 residues (R99, R101, R102 and D205) outside the pockets which have interactions of the tRNA moiety reported in the literature (Fig. 2A).Open in a separate windowFig. 2(A) Structural analysis of the AlbC (PDB ID, 3OQV). Enlarged view of the catalytically active pocket. The possible catalytic residues are shown in green (pocket-1) and orange (pocket-2) and basic residues on helix α4 are labelled in cyan. (B) Heatmap of the relative activity change in the cFL production levels of mutation of 14 amino acid residues in AlbC. WT residues are labeled with a black dashed rectangle. Dark blue boxes indicate that a specific mutant was not covered in randomly picked clones (activity assigned as −1).For SSM libraries, we adopted the “22c-trick” strategy to design degenerative primers, and 94 library clones were randomly picked for each residue library to reach a >98.6% probability of full library coverage of the 20 canonical amino acids.36 In addition to library variants, the WT and control strains were also included in the same 96-well plate. Overall, the abovementioned workflow (Fig. 1B) takes approximately 5 s for sample preparation and MS analysis for each mutant culture. Tentative ion peaks of cyclodipeptides were assigned based on theoretical m/z values (ESI Section 1) using the MetaboAnalyst webserver.37 For each ion peak, one-way analysis of variance (ANOVA) was used to evaluate statistical differences between the mean peak areas of a library variant and those of the control group. AlbC gene mutations were revealed by Sanger sequencing for all 14 SSM library members.To visualize the MALDI-ToF MS screening results, a heatmap was generated based on the cFL production of library members relative to that of WT (Fig. 2B), and we observed a general consistency between our results and literature data (Fig. 3A), respectively, which was similar to a previous study.26 Also, mutagenesis of basic residues (R99, R101 and R102) outside the pockets led to the decline of cFL production in the mutants, although R101 was more tolerant to mutations than R99 and R102. Furthermore, the substitution of D205 in the β6-α8 loop to alanine enhanced AlbC catalytic activity as observed previously.25 On the other hand, some previously unnoticed phenomena were observed. For example, in the D205 SSM library, relative ion intensities of cFL (m/z 261.17) in many mutants are higher than that of WT (Fig. S3). The top 3 mutants, D205M, D205K and D205R, were subsequently analyzed by LC-MS/MS. The results not only confirmed augmented synthesis of cFL, but also revealed increased production of other leucine-containing products including cYL, cLL and cML (Fig. 3B). Also interestingly, the discovery that most mutations of the T206 residue in the P2 pocket abolished the biosynthesis of the main product cFL was unprecedented (Fig. S3 and ESI Section 2). When examining MALDI mass spectra in detail, we noticed that although the T206F mutation greatly reduced cFL production, cFF and cFY products were largely not affected. Replacing the small threonine side chain with a bulky aromatic phenylalanine side chain could affect the activity of WT, because the phenyl ring structure increased the steric hindrance of T206F, thus impairing binding with its substrate and causing loss of its enzyme activity for cFL (Fig. 5A). These results suggested that T206 is also a key residue for the selectivity of the second Leu-tRNA substrate binding in the pocket. Unfortunately, we did not observe any new CDP molecules produced from the 14 SSM library mutants.Impact of AlbC mutations on the main product cFL
Target residueMutants and production levels relative to WT (literature)Mutants and production levels relative to WT (this study)
WT: cFL (set as 100%)WT: cFL (set as 100%)
Within binding pockets
L33/L185DL33Y/L185D: cFL (0)26L33Y: cFL (n.d.)a
L200L200N: cFL (<10%)26L200N: cFL (n.d.,a 5.6%)b
N159N159A: cFL (45%)25N159A: cFL (60%)a
Outside the pockets
R98R98A: cFL (20%)26
R99R99A: cFL (<10%)26R99A: cFL (n.d.)a
R98A/R99A: cFL (0)
R101/R102R101A/R102A: cFL (20%)26R101A: cFL (>50%)a
R102A: cFL (23%)a
D205D205A: cFL (>100%)25D205A: cFL (>100%)a
Open in a separate windowaCalculated from MALDI-TOF MS data.bCalculated from LC-MS/MS data.Open in a separate windowFig. 3Cyclodipeptide product analysis of AlbC mutants. (A) LC-MS characterization of the L200N mutant; (B) LC-MS characterization of select mutants in the D205 saturation mutagenesis library. Normalized production levels relative to WT (set as 1) are reported. Error bars represent the standard deviations of three biological replicates.Open in a separate windowFig. 5Structural illustration of the substrate binding pocket of the wildtype AlbC and the mutants F186L and T206. Note: Panel (A) Superimposition of the WT with T206F. Panels (B and C) Schematic representation of the hydrogen-bond and hydrophobic interaction between cFL·WT (B) and cFV·F186L (C). Dashed lines indicate hydrogen bond (yellow) interactions. The key amino acid residues of WT, F186L, and T206F are drawn as sticks in green, yellow and orange, respectively.The failure of isolating AlbC mutants with new substrate specificities from the SSM libraries highlights the limitation of semi-rational approaches that target manually selected residues. Therefore, we further turned to profile epPCR libraries that contain random mutations throughout the whole protein sequence. Under optimized conditions using the Agilent GeneMorph II Random Mutagenesis Kit, on average 2 nucleotides were introduced per gene variant so that most epPCR library members contain no more than one amino acid change. In total around 4500 independent clones were screened using the above workflow. The peak intensities at the theoretical m/z values of predicted cyclodipeptide ions in each MALDI mass spectrum were analyzed by the one-way ANOVA with Tukey''s multiple comparison test. A new peak (m/z 247) absent from WT that corresponds to the [M + H]+ ion of cFV was observed with three clones (Fig. 4A), all of which harbored the F186L mutation as revealed by Sanger sequencing results. Product identification was performed using LC-MS/MS and high-resolution (HR)-MS, and the retention time (5.7 min, Fig. 4B), exact mass (C14H19N2O2 [M + H]+m/z: 247.1441, Table S3), and MS/MS daughter ions23 (Fig. S4) of the new product were consistent with that of the chemically synthesized cFV standard. When comparing LC-MS/MS traces, we also noticed substantial reduction of the native main product cFL in the F186L variant relative to WT, and a slight enhancement in production of other products including cFM, cFF, and cFY, indicating a shift towards more bulky substrates with the F186L mutant (Fig. 4B). Another round of epPCR library screening was performed to further extend the AlbC substrate scope using the F186L variant as a parent, but unfortunately, no mutants were isolated to produce new cyclodipeptides.Open in a separate windowFig. 4(A) Box–whisker plots of 1152 samples at m/z 247.18 based on ANOVA analysis. Group 1, epRCR variants; 2, Rosetta(DE3)/pET28a-AlbC (WT); 3, Rosetta(DE3)/pET28a (control). Three clones containing F186L were labeled in red. ANOVA, analysis of variance. (B) LC-MS traces of WT and F186L mutant.Then, we investigated possible mechanisms underlying new substrate specificity of the F186L variant from structural aspects using computational modeling. In the substrate binding pocket, the cFL substrate formed three hydrogen bonds with two amino acid residues (N40 and E182) in the WT AlbC. The phenyl side chain of cFL forms a π–π stacking interaction with F186 (Fig. 5B and S5). However, when the residue at position 186 was mutated to leucine, this π–π stacking interaction was abolished. This leads to a change in the conformation of the Phe1 ring and thereby the hydrogen bond between Phe1 and N40 is abolished (Fig. 5C and S5), which ultimately results in outward movement of loop G33-S44 (Fig. S6). Therefore, the volume of the substrate binding cavity, which was measured to be 197 Å3, increased up to 288 Å3 when the residue at position 186 of WT was mutated to leucine (Table S4). Overall, it is possible that an increase in the production of cFV occurs precisely due to the larger volume of the substrate-binding cavity in F186L. Moreover, cFL and the new derivative cFV were separately docked into the binding pockets of the WT and F186L variant for 100 ns MD simulations (Fig. S5). The results showed that the binding energy of F186L was significantly increased when docking with the substrate cFV (l-lysine complexes (Fig. S5). Both WT and F186L reached the equilibrium state from an early stage.Binding energy of ligand for AlbC WT and F186L
Binding energya (kcal mol−1) of WTBinding energyb (kcal mol−1) of F186LBinding energyc (kcal mol−1) of F186L
ΔGVDWd−35.01−23.67−34.27
ΔGEte−31.95−8.70−29.37
ΔGpolarf48.7131.0847.60
ΔGapolarg−5.60−4.08−5.50
ΔGbindingh−23.85−5.37−21.54
Open in a separate windowaWT-cFL.bF186L-cFV.cF186L-cFL.dvan der Waals energy.eElectrostatic energy.fPolar-solvation energy.gNonpolar solvation energy.hΔGbinding = ΔGVDW + ΔGEt + ΔGpolar + ΔGapolar.In conclusion, we developed a robotic assay for directed evolution of AlbC, a model CDPS, using MALDI-ToF MS for label-free screening. Compared with conventional LC-MS (typically more than 5 min per sample), our HTS workflow represents a two-magnitude reduction of analytical time (5 s per sample). Contrary to previous reports that only study limited mutations of select residues, 14 SSM libraries were created and profiled for sequence-activity profiling, which not only confirmed the impact of known mutations (38,39 may also serve as label-free HTS assays, but further development is needed to address ion suppression issues caused by direct infusion of complex culture media into a mass spectrometer. On the other hand, the scarcity of positive hits in the single-residue SSM and epPCR libraries in this study confirmed AlbC as a difficult target for evolving new activities. Therefore, it is desirable to create and screen new AlbC mutant libraries that are comprehensive (i.e., deep mutational scanning40), combinatorial (i.e., combinatorial active site saturation test/iterative saturation mutagenesis, CAST/ISM41), or data-driven (i.e., machine learning-assisted directed evolution, MLDE42,43) in the future. Together, we envision that the label-free MS screening method should be generally applicable to engineering other enzymes with new activities.  相似文献   

13.
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   

14.
Fluorescent microscopy is a powerful tool for studying the cellular dynamics of biological systems. Small-molecule organic fluorophores are the most commonly used for live cell imaging; however, they often suffer from low solubility, limited photostability and variable targetability. Herein, we demonstrate that a tautomeric organic cage, OC1, has high cell permeability, photostability and selectivity towards the mitochondria. We further performed a structure–activity study to investigate the role of the keto–enol tautomerization, which affords strong and consistent fluorescence in dilute solutions through supramolecular self-assembly. Significantly, OC1 can passively diffuse through the cell membrane directly targeting the mitochondria without going through the endosomes or the lysosomes. We envisage that designing highly stable and biocompatible self-assembled fluorophores that can passively diffuse through the cell membrane while selectively targeting specific organelles will push the boundaries of fluorescent microscopy to visualize intricate cellular processes at the single molecule level in live samples.

In this article, we demonstrate the relatively unexplored potential of organic cages for use in targeted live cell imaging and highlight the importance of inter- and intramolecular interactions to stabilize and improve the performance of fluorophores.

Organic fluorophores are of great interest in many research fields especially live cell imaging, which allows for noninvasive observation and monitoring of biological processes. Small molecule fluorophores can be functionalized and tuned with high precision to tag a diverse variety of biological targets; however, limited water solubility, cell-permeability, photostability and targetability have halted the biomedical translation of many promising fluorescent molecules.1 Ultimately, new generations of fluorescent tags were designed to include new functionalities that can improve their physical properties and cell penetration such as long aliphatic side chains and cell penetrating peptides.2–6 In addition, various strategies have been employed to improve the photostability of these compounds by the addition of electron withdrawing groups, conformational changes, encapsulation or incorporation within host macrocycles.7–13Major efforts in the organic fluorophores development have been directed towards designing stable probes targeting the mitochondria as it plays a central role in the generation of adenosine triphosphate (ATP), central metabolism, and apoptosis.14,15 Successful examples of mitochondria targeting fluorophores include triphenyl phosphonium cations, heterocyclic aromatic cations, macrocyclic amphiphiles, BODIPY derivatives and mitochondria targeted peptides that have limited solubility.16–21 Commercially available dyes for mitochondrial imaging such as Mito Tracker Red have a high photobleaching tendency and can form covalent bonds with the mitochondrial respiratory chain complex I resulting in increased cytotoxicity. Consequently, various organic, inorganic and hybrid platforms were developed ranging from small nanoparticles to self-assembled frameworks on the quest for improved mitochondrial imaging agents.22–27Porous organic cages have received major attention over the last decade as they are easily prepared, soluble in a range of solvents and are intrinsically porous.28–31 They have shown excellent applicability in molecular recognition, sensing and hydrocarbon separation but their biomedical imaging applications have never been investigated.32–35 Recently, self-assembled organic imine stacks showed impressive electroluminescence and live cell imaging.36 Moreover, an intracellular targeted macrocyclic nanohoop showed fast cell uptake and two-photon live cell fluorescence imaging.37 In this work, a tautomeric organic cage (OC1) that shows unprecedented photostability for live-cell imaging with high biocompatibility, cell permeability and mitochondrial targetability is reported (Fig. 1). We hypothesize that the keto–enol tautomerism is playing a major role in promoting the stability and sustaining the fluorescence of this platform. Such effect has been previously reported in other organic dyes such as oxyluciferin and HDFL.38,39Open in a separate windowFig. 1Tautomeric forms of OC1 and chemical structures of OC2 and OC3.Investigating other organic cages with less hydroxyl groups (OC2) or replacing the hydroxyl groups with methoxy substituents (OC3) resulted in the loss of cell stability, retention and targetability (Fig. 1). The presence of the –OH groups in OC1 not only enables keto–enol tautomerization but also promotes H-bonding interactions, which ultimately improves solubility and cell permeability. We envision that understanding the nature of the intra- and intermolecular interactions in organic probes can lead to a profound enhancement in the photobleaching efficacy in addition to cell permeability. OC1 can effectively image the mitochondria over 3 h with high selectivity and without any loss of efficacy (no photo bleaching), which is comparable to the commercially used MitoTracker RedFM.OC1, a [2 + 3] cage, was synthesized following a reported procedure and its structure was confirmed by 1H Nuclear Magnetic Resonance (NMR) and TOF-Mass analysis (Fig. S1–S4).32 Dynamic light scattering (DLS) of OC1 revealed an average size of 30–50 nm in DMSO–H2O (2–5%) with a negative zeta potential of −5 mV (Fig. S5). Single crystal X-ray analysis revealed that OC1 crystallizes in a triclinic crystal system with chiral P1 space group and the asymmetric unit consists of two molecules of [2 + 3] cages (Fig. S6). Crystal structure analysis suggests that the imine cage crystallizes as the keto tautomeric form (Fig. 2a). This results in the formation of intramolecular N–H⋯O hydrogen bonding between the imine (N–H) hydrogen and the keto (C Created by potrace 1.16, written by Peter Selinger 2001-2019 O) oxygen (dC–OH actual = 1.36 Å; dC Created by potrace 1.16, written by Peter Selinger 2001-2019 O actual = 1.25–1.26 Å vs. dC Created by potrace 1.16, written by Peter Selinger 2001-2019 O for OC1 = 1.260 Å and dC Created by potrace 1.16, written by Peter Selinger 2001-2019 N actual = 1.25 Å vs. dC–NH for OC1 = 1.30–1.35 Å) (Fig. S6). Three different windows of each cage are connected with three different cages in the same plane through weak C–H⋯O hydrogen bonding (d = 2.414 Å, and θ = 60°), which creates isolated pores at the center of the cage (Fig. 2b). Furthermore, the packing diagram suggests the formation of one-dimensional infinite extrinsic pores along the b-axis (Fig. 2c and S6). OC1 is stable in aqueous solutions even at low pH as well as in cell media (Fig. S7–S10).Open in a separate windowFig. 2(a) SCXRD and crystal image of OC1. (b) OC1 self-assembled through hydrogen bonding interactions. (c) Packing of OC1 along b-axis.To test our original proposal that organic cages can act as effective imaging agents, we first investigated the UV-visible spectrum of OC1 in DMSO, which exhibited two characteristic peaks at λmax = 286 and 330 nm that can be assigned to π–π* and n–π* transitions, respectively (Fig. S11). Fluorescence emission spectra showed that OC1 exhibits two major peaks at around 411 and 497 nm upon excitation at 330–350 nm in DMSO, while excitation at 400 and 480 nm showed characteristic emission peaks at 503 and 538 nm, respectively (Fig. S12). This behavior can be explained by the presence of two emissive forms due to the keto–enol tautomerization. The quantum yield (QY) of OC1 in DMSO (0.25) and chloroform (0.4) was also tested. Molar extinction coefficient of OC1 was estimated as 46 806 M−1 cm−1 in DMSO (Fig. S13–S14). The quantum yield of OC1 in DMSO are about half that of in chloroform that is significantly matched with the reported intracellular targeted macrocyclic nanohoop structure.37 The fluorescent properties of OC1 were maintained at different pH values (4, 7 and 10) (Fig. S15). We hypothesize that this mode of assembly, as supported by SCXRD, promotes stability and prevents aggregation-induced quenching that is very common in small molecule dye fluorophores. The keto–enol tautomerism is boosting intra- and intermolecular hydrogen bonding interactions that can be contributing to the emissive characteristics of this system. In fact, the presence of the phenolic OH groups have been reported to have critical effects on the structural and spectroscopic properties of organic cages.31,40We then proceeded to test OC1 for cell imaging by first testing the cytotoxicity of OC1 in MCF-7 cells (Fig. S16a). The cells were treated with a solution of concentrations ranging from 3.9–500 μg ml−1 for 24 h. Cell death was measured using MTT assay demonstrating the excellent biocompatibility of OC1 with an IC50 of > 500 μg ml−1. MCF-7 cells were then either co-incubated with OC1 and LysoTracker Red or MitoTracker RedFM, the cells are then fixed with 4% formaldehyde (w/v) and imaged on a confocal laser-scanning microscope (CLSM). Interestingly, OC1 showed strong colocalization in the mitochondria with a Pearson''s correlation coefficient of 0.9751 (Fig. S16b and c). Real time imaging of MCF-7 cells at different time intervals after co-internalization of OC1 and MitoTracker showed a clear signal overlap on a subcellular level (Fig. 3 and MOV1), which ultimately supports that OC1 can selectively target the mitochondria.Open in a separate windowFig. 3Live cell imaging of MCF-7 cells incubated with 250 μg ml−1 of OC1 for 3 h showing the uptake and colocalization of OC1 (abs/em ∼495/519 nm) with the commercial probe MitoTracker RedFM (abs/em ∼581/644 nm) in the mitochondria (scale bar at 50 μm).These encouraging results piqued our interest to investigate other organic cages in order to have a better understanding of the mechanistic factors in play. As the tautomeric activity is well reported to affect the photophysical properties of organic compounds, we compared OC1 to an organic cage with less phenolic OHs (OC2) and an organic cage with only methoxy substituents (OC3). OC2 and OC3 were synthesized and characterized according to reported procedures (Fig. S17–S25).32,41 Absorbance, emission properties and stabilities of OC2 and OC3 in DMSO were also verified. OC2 showed one major characteristic emission peak at around 547 nm upon excitation at 350 and 440 nm. The quantum yield (QY) of OC2 in DMSO (6.7%) and chloroform (11.1%) was also tested (Fig. S13 and S14). OC3 showed emission at 494 and 537 nm upon excitation at 320 nm and 480 nm upon excitation at 420 nm (Fig. S26). The quantum yield (QY) and molar extinction coefficient of OC3 in DMSO is negligible. Cytotoxicity and biocompatibility studies of OC2 and OC3 showed a relatively higher cytotoxicity compared to OC1 (Fig. S27). As to targetability and photostability, OC2 and OC3 showed less selectivity with colocalization in both mitochondria and lysosome where the fluorescent signal slowly disappeared, which implies the OC2 and OC3 are less stable than OC1 (Fig. S28). Moreover, FACS data revealed that OC1 showed the highest uptake percentage in comparison to OC2, and OC3 with an uptake percentage of 86.4%, 78%, and 69%, respectively (Fig. S29). Based on this data, we can conclude that indeed the hydroxyl groups, and their prompted keto–enol tautomerism, are playing a major role in stabilizing the overall structure through hydrogen bonding while simultaneously improving photostability and targetability. As a control, we tested the aldehyde starting material, 1,3,5-triformylphloroglucinol, to verify that OC1 is being uptaken as a whole and not as the aldehyde building block or a possible decomposition product. The experiment clearly demonstrates that 1,3,5-triformylphloroglucinol as a fluorophore had no selectivity towards the mitochondria and is initially distributed in the cytoplasm and finally localizes in the lysosomes for degradation (Fig. S30).As for the uptake mechanism, we initially hypothesized that OC1 will be taken up by cells through endocytosis, which is the typical biological pathway of nanoparticles and nanoassemblies.42,43 Surprisingly, we discovered that OC1 can passively diffuse rather than being actively transported through the cell membrane. Passive transport is a type of membrane transport that does not require energy to move substances across cell membranes as it mainly relies on the second law of thermodynamics. We tested OC1 uptake in the presence of different endocytic inhibitors including chlorpromazine (CPZ), filipin (FIL), and amiloride to inhibit clathrin-mediated, caveolae mediated, and macropinocytotic endocytosis pathways, respectively.44 All cells were treated with the inhibitors for 30 min, followed by 3 h incubation with OC1. The cellular uptake data showed that there was no significant reduction in the uptake using the active transport inhibitors (Fig. 4a). Moreover, the passive transport was tested by incubating OC1 (125 μg ml−1) with MCF-7 for 3 h at different temperatures (4, 27, or 37 °C). The uptake was monitored by confocal microscopy (Fig. 4b) and quantified by FACS (Fig. 4c). Interestingly, no obvious reduction of uptake was observed in all the cases.Open in a separate windowFig. 4Cellular uptake mechanism of OC1. (a) Cell uptake of OC1 in the presence of endocytosis inhibitors to investigate endocytic uptake mechanism. (b) MCF-7 cell uptake CLSM of OC1 at 37 °C, 25 °C, and 4 °C, respectively (scale bar at 30 μm). (c) Quantification of cell fluorescence relative to 37 °C.After verifying the targetability and the uptake mechanism of OC1, we focused our attention on photostability. Since the commercially available mitochondrion tracker dyes normally have moderate photostability, it is critical to evaluate the photostability of OC1 in live cells. We first compared the photostability of OC1 with the commercially available MitoTracker RedFM. Since the excitation/emission wavelengths of OC1 (abs/em ∼495/519 nm) and MitoTracker RedFM (abs/em ∼581/644 nm) are distant from each other, we performed simultaneous scans for both dyes for 100 scans at 1.5 s per scan by incubating OC1 and MitoTracker RedFM with MCF-7 cells at 37 °C for 48 h. The results showed that at a laser power density of P ∼ 0.4132 W cm−2OC1 (ε = 46 806 M−1 cm−1) had higher retention and photostability than that of the MitoTracker RedFM where OC1 retained 54% of its initial fluorescence intensity while the fluorescence of MitoTracker RedFM dropped below 15% (Fig. 5a). Moreover, OC1 showed high photostability even at 72 h post-staining (Fig. 5b). Finally, it was noted that throughout the 72 h of incubation of OC1 with MCF-7, over three generations occurred in the cell culture and showed fluorescent labeling of the mitochondria suggesting that OC1 can be passed through to daughter cells upon cell division.Open in a separate windowFig. 5Intracellular retention and photostability of OC1. (a) Photostability of OC1 in MCF-7 cells co incubated with MitoTracker RedFM for 48 h. Plot of the sum intensity of each frame normalized to time 0 (frame 1) and plotted against time (seconds). Intensity data analysis was done using Python and plotted on Excel. (b) CLSM images (72 h post incubation) showing OC1 distribution while most of the MitoTracker RedFM signal was lost (Scale bar at 30 μm).  相似文献   

15.
A chelation-assisted oxidative addition of gold(i) into the C–C bond of biphenylene is reported here. The presence of a coordinating group (pyridine, phosphine) in the biphenylene unit enabled the use of readily available gold(i) halide precursors providing a new, straightforward entry towards cyclometalated (N^C^C)- and (P^C)-gold(iii) complexes. Our study, combining spectroscopic and crystallographic data with DFT calculations, showcases the importance of neighboring, weakly coordinating groups towards the successful activation of strained C–C bonds by gold.

Pyridine and phosphine directing groups promote the C–C activation of biphenylene by readily available gold(i) halides rendering a new entry to (N^C^C)- and (P^C)-gold(iii) species.

Activation of C–C bonds by transition metals is challenging given their inertness and ubiquitous presence alongside competing C–H bonds.1 Both the intrinsic steric hindrance as well as the highly directional character of the p orbitals involved in the σC–C bond impose a high kinetic barrier for this type of processes.2,3 Biphenylene, a stable antiaromatic system featuring two benzene rings connected via a four-membered cycle, has found widespread application in the study of C–C bond activation. Since the seminal report from Eisch et al. on the oxidative addition of a nickel(0) complex into the C–C bond of biphenylene,4 several other late transition metals have been successfully applied in this context.5 Interestingly, despite the general reluctance of gold(i) to undergo oxidative addition,6 its oxidative insertion into the C–C bond of biphenylene was demonstrated in two consecutive reports by the groups of Toste7a and Bourissou,7b respectively. The high energy barrier associated with the oxidation of gold could be overcome by the utilization of gold(i) precursors bearing ligands that exhibit either a strongly electron-donating character (e.g. IPr = [1,3-bis(2,6-diisopropylphenyl)imidazole-2-ylidene])7a or small bite angles (e.g. DPCb = diphosphino-carborane).7b,8 In line with these two approaches, more sophisticated bidentate (N^C)- and (P^N)-ligated gold(i) complexes have also been shown to aid the activation of biphenylene at ambient temperature (Scheme 1a).7c,dOpen in a separate windowScheme 1(a) Previous reports on oxidative addition of ligated gold(i) precursors onto biphenylene. (b) This work: pyridine- and phosphine-directed C–C bond activation of biphenylene by commercially available gold(i) halides.In this context, we hypothesized that the oxidative insertion of gold(i) into the C–C bond of biphenylene could be facilitated by the presence of a neighboring chelating group.9 This approach would not only circumvent the need for gold(i) precursors featuring strong σ-donor or highly tailored bidentate ligands but also offer a de novo entry towards interesting, less explored ligand templates. However, recent work by Breher and co-workers showcased the difficulty of achieving such a transformation.10Herein, we report the oxidative insertion of readily available gold(i) halide precursors into the C–C bond of biphenylene. The appendage of both pyridine and phosphine donors in close proximity to the σC–C bond bridging the two aromatic rings provides additional stabilization to the metal center and results in a de novo entry to cyclometalated (N^C^C)- and (P^C)gold(iii) complexes (Scheme 1b).Our study commenced with the preparation of 5-chloro-1-pyridino-biphenylene system 2via Pd-catalyzed Suzuki cross coupling reaction between 2-bromo-3-methylpyridine and 2-(5-chlorobiphenylen-1-yl)-4,4,5,5-tetramethyl-1,3,2-dioxaborolane 1 (Scheme 2).11 To our delight, the reaction of 2 with gold(i) iodide in toluene at 130 °C furnished complex κ3-(N^C^C)Au(iii)–I 3 in 60% yield.12,13 Complex 3 was isolated as yellow plate-type crystals from the reaction mixture and its molecular structure was unambiguously assigned by NMR spectroscopy, high-resolution mass spectrometry (HR-MS) and crystallographic analysis. Complex 3 exhibits the expected square-planar geometry around the metal center, with a Au–I bond length of 2.6558(3) Å.14 The choice of a neutral weakly bound gold(i)-iodide precursor is key for a successful reaction outcome: similar reactions in the presence of [(NHC)AuCl + AgSbF6] failed to deliver the desired biscyclometalation adducts, as reported by Breher et al. in ref. 10. The oxidative insertion of gold(i) iodide into the four-membered ring of pyridino-substituted biphenylene provides a novel and synthetically efficient entry to κ3-(N^C^C)gold(iii) halides. These species have recently found widespread application as precursors for the characterization of highly labile, catalytically relevant gold(iii) intermediates,15ad as well as for the preparation of highly efficient emitters in OLEDs.15eg Previous synthetic routes towards these attractive biscyclometalated gold(iii) systems involved microwave-assisted double C–H functionalization reactions that typically proceed with low to moderate yields.15aOpen in a separate windowScheme 2Synthesis of complex 3via oxidative addition of Au(i) into the C–C bond of pyridine-substituted biphenylene. X-ray structures of complex 3 with atoms drawn using 50% probability ellipsoids. Hydrogen atoms have been omitted for clarity. Additional selected bond distances [Å]: N–Au = 2.126(2), C1–Au = 1.973(2), C2–Au = 2.025(2), Au–I = 2.6558(3) and bond angles [deg]: N–Au–I = 99.25(6), N–Au–C1 = 79.82(9), C1–Au–C2 = 81.2(1), C2–Au–I = 99.73(8). For experimental details, see ESI.Encouraged by the successful results obtained with the pyridine-substituted biphenylene and considering the prominent use of phosphines in gold chemistry,6,16 we wondered whether the same reactivity would be observed for a P-containing system. To this end, both adamantyl- and tert-butyl-substituted phosphines were appended in C1 position of the biphenylene motif. Starting from 5-chlorobiphenylene-1-carbaldehyde 4, phosphine-substituted biphenylenes 5a and 5b could be accessed in 3 steps (aldehyde reduction to the corresponding alcohol, Appel reaction and nucleophilic displacement of the corresponding benzylic halide) in 64 and 57% overall yields, respectively.13 The reactions of 5a and 5b with commercially available gold(i) halides (Me2SAuCl and AuI) furnished the corresponding mononuclear complexes 7a–b and 8a–b, respectively (Scheme 3).13 All these complexes were fully characterized and the structures of 7a, 7b and 8a were unambiguously characterized by X-ray diffraction analysis.13 Interestingly, the nature of the halide has a clear effect on the chemical shift of the phosphine ligand so that a Δδ of ca. 5 ppm can be observed in the 31P NMR spectra of 7a–b (Au–Cl) compared to 8a–b (Au–I), the latter being the more deshielded. The Au–X bond length is also impacted, with a longer Au–I distance (2.5608(1) Å for 8a) compared to that measured in the Au–Cl analogue (2.2941(7) Å for 7a) (Δd = 0.27 Å).13Open in a separate windowScheme 3Synthesis and reactivity of complexes 7a–b, 8a–b, 9 and 10. X-ray structure of complexes 11b, 12 and 14 with atoms drawn using 50% probability ellipsoids. Hydrogen atoms have been omitted for clarity. For experimental details and X-ray structures see ESI.Despite numerous attempts to promote the C–C activation in these complexes,10,13 all reactions resulted in the formation of highly stable cationic species 11a–b and 12, which could be easily isolated from the reaction media. In the case of cationic mononuclear-gold(i) complexes 11, a ligand scrambling reaction in which the chloride ligand is replaced by a phosphine in the absence of a scavenger, a process previously described for gold(i) species, can be used to justify the reaction outcome.17 The formation of dinuclear gold complex 12 can be ascribed to the combination of a strong aurophilic interaction between the two gold centers (Au–Au = 2.8874(4) Å) and the stabilizing η2-coordination of the metal center to the aromatic ring of biphenylene. Similar η2-coordinated gold(i) complexes have been reported but, to the best of our knowledge, only as mononuclear species.18Taking into consideration the observed geometry of complexes 7a–b in the solid state,13 the facile formation of stable cationic species 11 and 12 and the lack of reactivity of the gold(i) iodides 8a–b, we hypothesized that the free rotation around the C–P bond was probably restricted, placing the gold(i) center away from the biphenylene system and thus preventing the desired oxidative insertion reaction. To overcome this problem, we set out to elongate the arm bearing the phosphine unit with an additional methylene group, introduced via a Wittig reaction from compound 4 to yield ligand 6, prepared in 4 steps in 27% overall yield. Coordination with Me2SAuCl and AuI resulted in gold(i) complexes 9 and 10, respectively (Scheme 3). The structure of 9 was unambiguously assigned by X-ray diffraction analysis and a similar environment around the metal center to that determined for complex 7a was observed for this complex.13With complexes 9 and 10 in hand, we explored their reactivity towards C–C activation of the four-membered ring of biphenylene.19 After chloride abstraction and upon heating at 100 °C for 5 hours, ring opening of the biphenylene system was observed for complex 9. Interestingly, formation of mono-cyclometalated adduct 13 was exclusively observed (the structure of 13 was confirmed by 1H, 13C, 31P, 19F, 11B and 2D NMR spectroscopy and HR-MS).13 The solvent appears to play a major role in this process, as performing the reaction in non-chlorinated solvents resulted in stable cationic complexes similar to 11.13,20,21 The presence of adventitious water is likely responsible for the formation of the monocyclometalated (P^C)gold(iii) complex 13 as when the reaction was carried out in C2H4Cl2 previously treated with D2O, the corresponding deuterated adduct 13-d could be detected in the reaction media. These results showcase the difficulties associated with the biscyclometalation for P-based complexes as well as the labile nature of the expected biscyclometalated adducts. Interestingly though, these processes can be seen as a de novo entry towards relatively underexplored (P^C)gold(iii) species.22The C–C activation was further confirmed by X-ray diffraction analysis of the phosphonium salt 14, which arise from the reductive elimination at the gold(iii) center in 13 upon exchange of the BF4 counter-anion with the weakly coordinating sodium tetrakis[3,5-bis(trifluoromethyl)phenyl]borate (NaBArF).13,23 The phosphorus atom is four-coordinate, with weak bonding observed to the distant counter-anion and a distorted tetrahedral geometry (C1–P–C2 = 95.05(17), C2–P–C3 = 112.1(1), C3–P–C4 = 116.6(1), C4–P–C1 = 107.4(2) deg). These results represent the third example in which the C(sp2)–P bond reductive elimination at gold(iii) has been reported.24Further, it is important to note that, in contrast to the reactivity observed for the pyridine-substituted biphenylene, neither P-coordinated gold(i) iodo complexes 8a, 8b nor 10 reacted to give cyclometalated products despite prolonged heating, which highlights the need for highly reactive cationized gold(i) species to undergo oxidative addition when phosphine ligands are flanking the C–C bond.13To get a deeper understanding on the observed differences in reactivity for the N- vs. P-based directing groups, ground- and transition-state structures for the oxidative insertion of gold(i) halides in C1-substituted biphenylenes were computed by DFT calculations. The reactions of Py-substituted 2 with AuI to give 3 (I) and those of P-substituted 7a (II) and 9 (III) featuring the cationization of the gold(i) species were chosen as models for comparative purposes with the experimental conditions (Fig. 1 and S1–S10 in the ESI).25–27 The computed activation energies for the three processes are in good agreement with the experimental data. The pyridine-substituted biphenylene I exhibits the lowest activation barrier for the oxidative insertion process (ΔG = 34.4 kcal mol−1). The reaction on the phosphine-substituted derivatives II and III proved to be, after cationization of the corresponding gold(i) halide complexes (II-BF4, III-BF4) higher in energy (ΔG = 39.6 and 46.3 kcal mol−1 respectively), although the obtained values do not rule out the feasibility of the C–C activation process. The transition state between I and I′ exhibits several interesting geometrical features: (a) the biphenylene is significantly bent, (b) the cleavage of the C–C bond is well advanced (dC–C = 1.898 Å in TSIvs. dC–C = 1.504 Å in I), and (c) the two C and the I atoms form a Y-shape around gold with minimal coordination from the pyridine (dN–Au = 2.742 Å in TSIvs. dN–Au = 2.093 Å in I and 2.157 Å in I′, respectively). The transition-state structures found for the P-based ligands (TSII and TSIII) also show an elongation of the C–C bond and display a bent biphenylene. However, much shorter P–Au distances (dP–Au = 2.330 Å for TSII and 2.314 Å for TSIII) can be observed compared to the pyridine-based system, as expected due to the steric and electronic differences between these two coordinating groups. Analogously, longer C–Au distances were also found for the P-based systems (dC1–Au = 2.152 Å for TSIvs. 2.235 Å and 2.204 Å for TSII and TSIII; dC2–Au = 2.143 Å for TSIvs. 2.219 Å and 2.162 Å for TSII and TSIII), with a larger deviation of square planarity for Au in TSIII compared to TSII.28,29 These results suggest that, provided the appropriate distance to the C–C bond is in place, the strong coordination of phosphorous to the gold(i) center does not prevent the C–C activation of biphenylene but other reactions (i.e. formation of diphosphine gold(i) cationic species, protodemetalation) can outcompete the expected biscyclometalation process. In contrast, a weaker donor such as pyridine offers a suitable balance bringing the gold in close proximity to the C–C bond and enables both the oxidative cleavage as well as the formation of the double metalation product.Open in a separate windowFig. 1Energy profile (ΔG and ΔG in kcal mol−1), optimized structures, transition states computed at the IEFPCM (toluene/1,2-dichloroethane)-B3PW91/DEF2QZVPP(Au,I)/6-31++G(d,p)(other atoms) level of theory for the C–C activation of biphenylene with gold(i) iodide from I and gold(i) cationic from II and III. Computed structures of the transition states (TSI, TSII and TSIII) and table summarizing relevant distances.  相似文献   

16.
A molecular keypad lock that displays photodynamic activity when exposed to glutathione (GSH), esterase and light in the given order, is fabricated and its efficacy in drug resistant MCF7 cancer cells is investigated. The first two inputs are common drug resistant tumor markers. GSH reacts with the agent and shifts the absorption wavelength. Esterase separates the quencher from the structure, further activating the agent. After these sequential exposures, the molecular keypad lock is exposed to light and produces cytotoxic singlet oxygen. Among many possible combinations, only one ‘key’ can activate the agent, and initiate a photodynamic response. Paclitaxel resistant MCF7 cells are selectively killed. This work presents the first ever biological application of small molecular keypad locks.

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

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

17.
Copper-catalyzed electrochemical direct chalcogenations of o-carboranes was established at room temperature. Thereby, a series of cage C-sulfenylated and C-selenylated o-carboranes anchored with valuable functional groups was accessed with high levels of position- and chemo-selectivity control. The cupraelectrocatalysis provided efficient means to activate otherwise inert cage C–H bonds for the late-stage diversification of o-carboranes.

Copper-catalyzed electrochemical cage C–H chalcogenation of o-carboranes has been realized to enable the synthesis of various cage C-sulfenylated and C-selenylated o-carboranes.

Carboranes are polyhedral molecular boron–carbon clusters, which display unique properties, such as a boron enriched content, icosahedron geometry and three-dimensional electronic delocalization.1 These features render carboranes as valuable building blocks for applications to optoelectronics,2 as nanomaterials, in supramolecular design,3 organometallic coordination chemistry,4 and boron neutron capture therapy (BNCT) agents.5 As a consequence, considerable progress has been witnessed in transition metal-catalyzed regioselective cage B–H functionalization of o-carboranes6 and different functional motifs have been incorporated into the cage boron vertices.7–10 However, progress in this research arena continues to be considerably limited by the shortage of robust and efficient methods to access carborane-functionalized molecules. While C–S bonds are important structural motifs in various biologically active molecules and functional materials,11 strategies for the assembly of chalcogen-substituted carboranes continue to be scarce. A major challenge is hence represented by the strong coordination abilities of thiols to most transition metals, which often lead to catalyst deactivation.12 While copper-catalyzed B(4,5)–H disulfenylation of o-carboranes was achieved,7e elevated reaction temperature was required, and 8-aminoquinoline was necessary as bidentate directing group. The bidentate directing group13 needs to be installed and removed, which jeopardizes the overall efficacy. Likewise, an organometallic strategy was recently devised for cysteine borylation with a stoichiometric platinum(ii)-based carboranes.14 Meanwhile, oxidative cage B/C–H functionalizations largely call for noble transition metal catalysts15 and stoichiometric amounts of chemical oxidants, such as expensive silver(i) salts.16In recent years, electricity has been identified as an increasingly viable, sustainable redox equivalent for environmentally-benign molecular synthesis.17,18 While significant advances have been realized by the merger of electrocatalysis with organometallic bond activation,19 electrochemical carborane functionalizations continue unfortunately to be underdevelopment. In sharp contrast, we have now devised a strategy for unprecedented copper-catalyzed electrochemical cage C–H chalcogenations of o-carboranes in a dehydrogenative manner, assembling a variety of C-sulfenylated and C-selenylated o-carboranes (Fig. 1a). It is noteworthy that our electrochemical cage C–S/Se modification approach is devoid of chemical oxidants, and does not need any directing groups, operative at room temperature.Open in a separate windowFig. 1Electrochemical diversification of o-carboranes and optimization of reaction conditions. aReaction conditions: procedure A: 1a (0.10 mmol), 2a (0.3 mmol), CuOAc (15 mol%), 2-PhPy (15 mol%), LiOtBu (0.2 mmol), TBAI (2.0 equiv.), solvent (3 mL), platinum cathode (10 mm × 15 mm × 0.25 mm), graphite felt (GF) anode (10 mm × 15 mm × 6 mm), 2 mA, under air, r.t., 16 h. bYield was determined by 1H NMR with CH2Br2 as the internal standard. cIsolated yields in parenthesis. dKI (1.0 equiv.) as additive. eProcedure B: 2 (0.3 mmol), LiOtBu (0.2 mmol), TBAI (2.0 equiv.), solvent (3.0 mL), 2 mA, r.t., 3 h, then adding 1a (0.10 mmol), 2-PhPy (15 mol%), CuOAc (15 mol%), 2 mA, rt, 16 h. f2b (0.3 mmol), LiOtBu (0.2 mmol), KI (1.0 equiv.), TBAI (2.0 equiv.), solvent (3.0 mL), 2 mA, r.t., 3 h, then adding 1a (0.10 mmol), 2-PhPy (15 mol%), CuOAc (15 mol%), r.t., 16 h. TBAI = tetrabutylammonium iodide, TBAPF6 = tetrabutylammonium hexafluorophosphate. DCE = 1,2-dichloroethane, THF = tetrahydrofuran.We commenced our studies by probing various reaction conditions for the envisioned copper-catalyzed cage C–H thiolation of o-carborane in an operationally simple undivided cell setup equipped with a GF (graphite felt) anode and a Pt cathode (Fig. 1b and Table S1). After extensive experimentation, we observed that the thiolation of substrate 1 proceeded efficiently with catalytic amounts of CuOAc and 2-phenylpyridine, albeit in the presence of 2 equivalents LiOtBu as the base, and 2 equivalents n-Bu4NI as the electrolyte at room temperature under a constant current of 2 mA (entry 1). The yield was reduced when other copper sources or additives were used (entries 2–5). Surprisingly, n-Bu4NPF6 as the electrolyte failed to facilitate the carborane modification, indicating that n-Bu4NI operates not only as electrolyte, but also as a redox mediator (entry 6). Altering the stoichiometry of the electrolyte or using KI did not improve the performance (entries 7–8). Product formation was not observed, when the reaction was conducted with DCE as the solvent, while CH3CN resulted in a drop of the catalytic performance (entries 9–10). Control experiments confirmed the essential role of the electricity and the catalyst (entries 11–12), while a sequential procedure was found to be beneficial (entries 13–15).With the optimized reaction conditions in hand, we explored the versatility of the cage C–H thiolation of o-carborane 1a with different thiols 2 (Scheme 1). Electron-rich as well as electron-deficient substituents on the arenes were found to be amenable to the electrocatalyzed C–H activation, providing the corresponding thiolation products 3aa–3ao in good to excellent yields. Thereby, a variety of synthetically useful functional groups, such as fluoro (3ae, 3am), chloro (3af, 3ak, 3an) and bromo (3ag, 3al), were fully tolerated, which should prove instrumental for further late-stage manipulations. Various disubstituted aromatic and heterocyclic thiols afforded the corresponding cage C–S modified products 3ap–3as. Notably, aliphatic thiols efficiently underwent the electrochemical transformation to provide the corresponding cage alkylthiolated products 3at–3au. Notably, the halogen-containing thiols (2e–2f, 2k–2n and 2q) reacted selectively with o-carboranes to deliver the desired products without halide coupling byproducts being observed. The connectivity of the products 3aa, 3am and 3ao was unambiguously verified by X-ray single crystal diffraction analysis.22Open in a separate windowScheme 1Electrochemical C–H thiolation of o-carborane 1a. (a) Procedure B. (b) KI (1 equiv.). (c) Cul as the catalyst.Encouraged by the efficiency of the cupraelectro-oxidative cage C–H thiolation, we became intrigued to explore the chalcogenantion of differently-decorated o-carboranes 1 (Scheme 2). Electronically diverse carboranes 1 served as competent coupling partners, giving the corresponding thiolation products 4bo–4do with high levels of efficacy in position-selective manner. The strategy was not restricted to phenyl-substituted o-carboranes. Indeed, substrates bearing benzyl and even alkyl groups also performed well to deliver the desired products 4eo–4ga. It is noteworthy that the C–H activation approach was also compatible with selenols to give the o-carboranes 4av–4fv. The molecular structures of the carborane 4br and 4av were unambiguously verified by single-crystal X-ray diffraction.22Open in a separate windowScheme 2Electrochemical cage C–H chalcogenation of o-carboranes. (a) Procedure B. (b) KI (1 equiv.).Scaffold functionalization of the thus obtained carborane 3ag provided the alkynylated derivative 5a and amine 5b (Scheme 3), giving access to carborane-based host materials of relevant to phosphorescent organic light-emitting diodes.20Open in a separate windowScheme 3Late-stage diversification.Next, we became attracted to delineating the mode of the cupraelectro-catalyzed cage C–H chalcogenation. To this end, control experiments were performed (Scheme 4a). First, electrocatalysis in the presence of TEMPO or Ph2C Created by potrace 1.16, written by Peter Selinger 2001-2019 CH2 gave the desired product 3aa. EPR studies of thiol 2a, LiOtBu and THF under the electrochemical conditions showed a small radical signal, which might be attributed to a thiol radical.21 Second, the cupraelectrocatalysis occurred efficiently in the dark. Third, detailed cyclovoltammetric analysis of the thiol and iodide mediator (Scheme 4b and ESI)21 revealed an irreversible oxidation of the thiol anion at Ep = −0.62 V vs. Ag/Ag+ and two oxidation events for the iodide, including an irreversible oxidation at Ep = 0.12 V vs. Ag/Ag+ and a reversible oxidation at Ep = 0.44 V vs. Ag/Ag+, which is in good agreement with the literature reported iodide oxidation potentials,18c,d and is suggestive of the preferential oxidation of the iodide as a redox mediator. In this context, the use of n-Bu4NI as a redox mediator to achieve copper-catalyzed electrochemical arene C–H aminations had been documented.18d Furthermore, we calculated the redox potential of complex C by means of DFT calculations at the PW6B95-D4/def2-TZVP + SMD(MeCN)//TPSS-D3BJ/def2-SVP level of theory.21 These studies revealed a calculated oxidation half-wave potential for complex C is Eo,calc1/2 = −0.08 V vs. SCE. Hence, iodide is a competent redox mediator to achieve the transformation from complex C to complex D. Analysis of non-covalent interactions21 in complex C (Fig. 2) show the presence of a weak stabilization interaction between the chalcogen''s anisole group and the 2-phenylpyridine. In contrast, in complex D these interactions were found more relevant between the o-carborane phenyl group and the chalcogen aromatic motif.Open in a separate windowFig. 2Non-covalent interaction plots for the complexes C and D. Strong attractive interactions are shown in blue, weak attractive interactions are given in green, while red corresponds to repulsive interactions. Ar = 4-MeOC6H4.Open in a separate windowScheme 4Control experiments and cyclic voltammograms.On the basis of the aforementioned findings,18 a plausible reaction mechanism is proposed in Scheme 5, which commences with an anodic single electron-transfer (SET) oxidation of the thiol anion E to form the sulfur-centered radical F. Subsequently, the copper(i) species A reacts with the sulfur radical F to deliver copper(ii) complex B, which next reacts with o-carborane 1 in the presence of LiOtBu to generate a copper(ii)-o-carborane complex C. Thereafter, the complex C is oxidized by the anodically generated redox mediator I2 to furnish the copper(iii) species D,18d which subsequently undergoes reductive elimination, affording the final product and regenerating the catalytically active complex A. Alternatively, the direct oxidation of copper(ii) complex C by electricity to generate copper(iii) species D can not be excluded at this stage.18a,bOpen in a separate windowScheme 5Proposed reaction mechanism.In conclusion, a sustainable electrocatalytic C–H chalcogenation of o-carboranes with thiols and selenols was realized at room temperature by earth abundant copper catalysis. The C–H activation was characterized by mild reaction conditions and high functional group tolerance, leading to the facile assembly of various o-carboranes. Thereby, a transformative platform for the design of cage C–S and C–Se o-carboranes was established that avoids chemical oxidants by environmentally-sound electricity in the absence of directing groups. A plausible mechanism of paired electrolysis was established by detailed mechanistic studies.  相似文献   

18.
Odd-electron bonds have unique electronic structures and are often encountered as transiently stable, homonuclear species. In this study, a pair of copper complexes supported by Group 13 metalloligands, M[N((o-C6H4)NCH2PiPr2)3] (M = Al or Ga), featuring two-center/one-electron (2c/1e) σ-bonds were synthesized by one-electron reduction of the corresponding Cu(i) ⇢ M(III) counterparts. The copper bimetallic complexes were investigated by X-ray diffraction, cyclic voltammetry, electron paramagnetic spectroscopy, and density functional theory calculations. The combined experimental and theoretical data corroborate that the unpaired spin is delocalized across Cu, M, and ancillary atoms, and the singly occupied molecular orbital (SOMO) corresponds to a σ-(Cu–M) bond involving the Cu 4pz and M ns/npz atomic orbitals. Collectively, the data suggest the covalent nature of these interactions, which represent the first examples of odd-electron σ-bonds for the heavier Group 13 elements Al and Ga.

Hanging on by a thread. Formally zerovalent copper complexes with an Al(iii) or Ga(iii) support were investigated. The combined experimental and theoretical data corroborate the presence of an odd-electron σ-bond between Cu and the Group 13 center.

Odd-electron σ-bonds, where the electrons are delocalized between two atoms, can occur as two-center/one-electron (2c/1e) or two-center/three-electron (2c/3e) interactions. Proposed by Pauling in 1931,1 odd-electron σ-bonds have garnered attention because of their fundamental importance to chemical bonding and their relationship to radical species generated during oxidative stress in biological systems.2–14 Examples of compounds exhibiting odd-electron bonding are typically homonuclear (like H2+, He2+, and alkali metal dimers) and transiently stable, limiting them to spectroscopic characterization.1,11,15–18The first solid-state structure of a formally one-electron σ-bond was a tetraphosphabenzene species (Fig. 1a) which was formed by the coupling of two diphosphirenyl radicals.19 Following this discovery, the formation of discrete 2c/1e σ-bonds, where the odd-electron is delocalized between two homonuclear main group centers, was reported for B·B and then extended to P·P.8,17,20 Of note, the first solid-state structure of a B·B compound was reported in only 2014 (Fig. 1b).21 Examples of 2c/1e σ-bonds between the heavier Group 13 congeners are even more lacking because of the greater propensity for their unpaired spins to couple, forming larger more stable clusters.8 To our knowledge, there are only three structurally characterized examples of odd-electron bonds for the heavy Group 13 atoms,22 and these examples are all homonuclear π-radicals (Fig. 1c).23–26Open in a separate windowFig. 1Select examples of structurally characterized molecules (a–d) featuring odd-electron bonds.Heteronuclear odd-electron σ-bonds are also rare. The Cu(TPB) complex, where TPB is a trisphosphinoborane, is the single structural example of a 2c/1e bond between heteroatoms (Fig. 1d).27 The authors described the bonding as Cu·B, where the unpaired electron is heavily polarized toward B. A theoretical study predicted that such a bond would also exist between Cu and Al, but no heavier analogues of Cu(TPB) have been synthesized to date.28 Furthermore, the heavier Group 13 elements by virtue of their lower electronegativity compared to B should facilitate greater covalent interactions with the Cu center.Hence, we sought to target formally zerovalent Cu complexes supported by Al(III) or Ga(III) as an extension of the previously reported isoelectronic nickelate species and Cu(TPB).29 Herein, we describe the synthesis, structure, spectroscopic characterization, and DFT calculations of cationic [CuML]+ complexes (L = [N((o-C6H4)NCH2PiPr2)3]3−; M = Al and Ga) as well as their one-electron reduced metalloradical counterparts that feature discrete 2c/1e bonds.  相似文献   

19.
In any drug discovery effort, the identification of hits for further optimisation is of crucial importance. For peptide therapeutics, display technologies such as mRNA display have emerged as powerful methodologies to identify these desired de novo hit ligands against targets of interest. The diverse peptide libraries are genetically encoded in these technologies, allowing for next-generation sequencing to be used to efficiently identify the binding ligands. Despite the vast datasets that can be generated, current downstream methodologies, however, are limited by low throughput validation processes, including hit prioritisation, peptide synthesis, biochemical and biophysical assays. In this work we report a highly efficient strategy that combines bioinformatic analysis with state-of-the-art high throughput peptide synthesis to identify nanomolar cyclic peptide (CP) ligands of the human glucose-dependent insulinotropic peptide receptor (hGIP-R). Furthermore, our workflow is able to discriminate between functional and remote binding non-functional ligands. Efficient structure–activity relationship analysis (SAR) combined with advanced in silico structural studies allow deduction of a thorough and holistic binding model which informs further chemical optimisation, including efficient half-life extension. We report the identification and design of the first de novo, GIP-competitive, incretin receptor family-selective CPs, which exhibit an in vivo half-life up to 10.7 h in rats. The workflow should be generally applicable to any selection target, improving and accelerating hit identification, validation, characterisation, and prioritisation for therapeutic development.

mRNA display generates vast datasets of protein binders. Bioinformatic clustering of the sequences combined with high throughput synthesis and analysis methods allow efficient prioritisation of hits for in vivo experiments.

In recent years, peptides have emerged as powerful therapeutic agents1 and the most prominent examples can match antibodies in selectivity and potency. This development has been greatly aided by breakthroughs in half life extension technologies by formulation or chemical modification. A key advantage of engineered peptides is their ease of production, and while antibodies are generally considered to be too large for efficient oral uptake in therapeutically relevant amounts, the small size of peptides allows the potential for oral dosing with significant progress made in recent years.2,3 Despite these advantages, most drug development projects heavily rely on antibodies, most likely due to the ease of de novo antibody discovery by evolutionarily optimised methods. Modern in vivo or in vitro display technologies, such as phage display and mRNA display, allow for the identification of potent and specific peptide ligands through iterative screening of peptide pools containing trillions of randomised sequences.4,5 These technologies have the potential to yield a plethora of de novo peptides which could be engineered into powerful therapeutics. However, these techniques are surprisingly scarce in clinical pipelines (compared to antibodies or small molecules) despite being known for decades and most peptides in late-stage discovery are derivatives of naturally occurring molecules.1 Challenges in hit prioritisation and time-consuming validation could at least partly be responsible for this observation. While classical approaches would rely on Sanger sequencing of the most abundant clones (yielding very few hits), the advent of next generation sequencing (NGS) has revolutionized the information content available from de novo peptide screens.4 In the majority of studies, however, this information is unused and only a handful of the most abundant sequences are followed up by chemical synthesis for testing of biological activity.5 We rationalized that powerful bioinformatic data mining methods would best utilise the vast information contained in peptide display datasets. To prove this, we combined chemical synthesis and high throughput binding analysis on data from mRNA display. We show that these methods identify more diverse series of hits with different modes of binding as starting points for peptide drug discovery efforts, especially when a functional binder to a specific binding pocket is desired (Scheme 1).Open in a separate windowScheme 1Workflow of peptide hit identification, prioritisation, and validation.We chose the type 2 GPCR glucose-dependent insulinotropic peptide receptor (GIP-R) as a challenging and clinically relevant test case.6 Its natural ligand GIP and the close relative glucagon-like peptide 1 (GLP-1) are incretins, i.e., hormones which are secreted after oral nutrient uptake and which augment glucose-dependent insulin release from pancreatic beta cells.7 GLP-1 receptor agonists have been reported to not only have effects in metabolic disease through increasing insulin release, but also promote satiety by delaying gastric emptying, and are being investigated for their protective effects in obesity,8,9 heart disease,10 and diabetic kidney disease.11 Conversely, the biology of GIP is less clearly understood, and both GIP-R agonists12 as well as antagonists13,14 are under investigation as potential therapeutics in the fields of type 2 diabetes and obesity. Previous approaches to target the GIP-R thus far have made use of analogues of GIP itself,15,16 or GIP R specific antibodies14,17–20 and peptide display work was based on dual GIP-GLP-1 analogue libraries.14,19,20 As such, the GIP-R represented an ideal target for our mRNA based workflow as we strived to identify potent, ligand competitive, and subfamily selective de novo binders, which could provide valuable tools to understand GIP-R biology and serve as starting points for drug discovery efforts.We employed mRNA display to identify cyclic peptide ligands towards the biotinylated extracellular domain (ECD) of the human GIP receptor (residues 22–138 of hGIP-R), which has been shown to retain binding to GIP.21 Initially, we established conditions for efficient cyclisation by disulphide bond formation for representative library peptides (Fig. S3), before conducting iterative rounds of mRNA display5,22 using a nucleotide library encoding two conserved cysteines for macrocyclization flanking a region of 4 to 12 random amino acid sequence. Negative selections were performed against biotin-loaded streptavidin-functionalised magnetic beads, followed by positive selection against bead-immobilised hGIP-R ECD. Sufficient library enrichment was obtained for hGIP-R after five rounds of selections (percentage recovery relative to input > 0.1%) (Fig. S4a). NGS of enriched output cDNA was carried out and peptides were clustered by similarity for each selection round (R1–R5). All sequences exceeding six reads in R5 (the top 3160 sequences) were compared against each other using pairwise local alignments in order to derive a complete distance matrix. Single-linkage hierarchical clustering was then performed on the distance matrix, choosing a threshold for cutting the relative dendrogram into clusters based on a statistical optimality criterion.23 For this dataset, the sequence similarity threshold for assigning a sequence to a cluster was set at 0.38, and the fewest number of members that a cluster was allowed to contain was 20. This generated 13 clusters (assigned letters A through M), where the largest cluster, A, contained 1880 sequences, while the two least populated clusters L and M contained 21 sequences each, and the unclustered sequences were assigned to a “noise” cluster (termed cluster 0), encompassing 263 sequences (Fig. 1).Open in a separate windowFig. 1High throughput pairwise clustering analysis of 3160 Round 5 output sequences selected against hGIP-R ECD. Lighter colour indicates closer similarity between individual sequences (ESI Data 1).In the next step, we selected peptides from each cluster (A-M, 0) based on sequence diversity and abundance in R5 for parallel solid-phase peptide synthesis (SPPS) in a 96 well format. The N-terminus was capped with an acetyl group (Ac) to mimic the fMet in mRNA-display and a C-terminal FLAG tag was added to each peptide to ensure peptide solubility in buffer by serving as a source of negative charge, analogous to C-terminal mRNA that is present during the selection screens.24 Peptides were assigned identifiers with cluster ID and their rank order based on their abundance in R5 (e.g. B_5). Following synthesis and cleavage, the peptides were macrocyclized via disulphide bond formation in 20% DMSO in buffer.25 Peptide identity, purity, and complete macrocyclization was confirmed by UPLC-MS and peptide concentrations were determined by UPLC-CAD (ESI Data 3). For the initial high throughput characterisation, peptides were used without further purification (average purity 50–70%). To establish the binding properties, we performed high-throughput single-concentration biolayer interferometry (BLI) measurements, using biotinylated hGIP-R ECD immobilised on streptavidin-functionalised biosensors. Select peptides with favourable binding potencies were resynthesized, purified and their Kd determined by multiple concentration BLI measurements (Fig. 2). The results were generally in good agreement with the single concentration data.Open in a separate windowFig. 2(a and b) Single concentration hGIP-R ECD binding Kd values and dissociation rates for the most abundant peptide clusters, and corresponding abundances/reads of peptide sequences in the final round of selection against GIP-R ECD. All peptides contain an N-terminal Ac group and C-terminal FLAG tag and are cyclised as disulphides. (c) Radiolabelled [125I]-GIP displacement from hGIPR#5/BHK Creluc 2P cells by cyclic peptides at a single concentration of 1000 nM CPs (x axis), and 100 nM CPs (y axis) (CPM = counts per minute). All peptides were tested crude without prior purification (ESI Data 2).We identified several peptides (>50% of our panel) with nanomolar binding affinities toward the hGIP-R from multiple sequence clusters, particularly from the two most abundant clusters A and B (Fig. 2a and b). However, potent binders were also identified in less abundant sequence families. In particular, peptide M_46 was the second most potent binder amongst the resynthesized peptide panel, while only accumulating 0.2% of the reads compared to the most abundant peptide A_0. This highlights the benefit of following up on peptide hits independent of their ranking based on sequencing reads. Encouragingly, weak-to-no binding was observed from peptide members of cluster 0 (Fig. S8), corresponding to bioinformatic ‘noise’ in the NGS data, which likely consists of singleton sequences and possible sequencing errors. The FLAG tag itself did not bind to hGIP-R ECD. For members of clusters containing an internal cysteine, e.g. peptide C_24, we generated single cysteine-to-methionine mutants for each site (Met2, Met8, and Met13) and deduced the disulphide pattern from the binding data (i.e. a bond between Cys2 or Cys8 for C_24, see Fig. S8). In the spirit of the high throughput nature of the workflow, we did not follow up on peptides which failed during parallel synthesis, as the goal was the identification of one hit sequence. As a consequence, we suggest that the pursuit of peptides with complex folds or unpaired cysteines is only indicated if no functional hits can be identified otherwise.The non-biased selection process means that high affinity CPs can be generated against any sites on hGIP-R ECD. We hypothesized that each cluster represents peptides which bind at a specific target binding site, with the possibility of different clusters having overlapping or distinct binding sites on the receptor. In this study, we were interested in identifying inhibitory peptide families which are able to either compete or disrupt the natural ligand, whether through direct competition or allosteric binding. To investigate this, we selected representative CPs from each cluster and performed competitive displacement assays with 125I-GIP in BHK CreLuc 2P cells stably expressing hGIP-R. Notably, this assay shows (competitive) binding to the full length GIP-R, whereas our selection and initial screen had been performed against the ECD. As shown in Fig. 2c, only peptides from clusters B and M were found to displace 125I-GIP from hGIP-R (as measured at peptide concentrations of 100 nM and 1000 nM). Interestingly, cluster B contains a LWPF motif at the C-terminus, while cluster M has a related LPWF motif at the N-terminus, indicating a potentially conserved binding site. None of the members of other clusters were found to displace 125I-GIP, even those which were determined to have nanomolar binding affinities to the hGIP R ECD by BLI, including the most abundant peptide family of the selection (cluster A), which covered over 50% of the whole dataset. This highlights the value of the clustering approach, as enrichment of the sequences during mRNA display is driven only by binding affinities of encoded CPs to hGIP-R ECD, but our subsequent clustering analyses served to reveal different functionalities of these CPs. Thus, the high-throughput prioritisation approach is likely to have a higher chance of identifying functional ligands, rather than those prioritised based on observed amounts of NGS reads.Encouraged by these results, we focussed our subsequent work on cluster B, namely the structurally most diverse members B_3, B_5, and B_68. B_68 had < 20% purity in the 96 well synthesis and did not reveal any binding in the initial high throughput binding assay but was included in the follow up. Further analysis of the sequence list showed another structurally diverse member of cluster B, B_1275, which we hypothesized would also exhibit potent binding and GIP competition, despite the low levels of sequencing reads. These four sequences were scaled up by traditional SPPS and further investigated in multi-concentration BLI and competitive binding assays. All purified peptides showed nanomolar binding and competition of the 125I-GIP from BHK cells stably expressing hGIP-R (Fig. 4). Furthermore, none of these ligands were found to displace radiolabelled GLP-1 from GLP-1R, nor glucagon from GCG-R (Fig. S11). The parent CP sequences also do not bear any significant sequence homology with native GIP, nor any known interactors of hGIP-R, suggesting that these are the first reported examples of de novo incretin receptor selective and competitive peptides for hGIP-R.Open in a separate windowFig. 4(a) BLI binding Kd values and radiolabelled [125I]-GIP displacement IC50 data for purified peptides. Multiple concentrations of CPs were used to determine these values. (*) uncertain values due to limited solubility of CP. (b) Stability of selected peptides upon incubation with human plasma at 37 °C. (c) In vivo plasma exposure levels of selected peptides upon i.v. dosing to rats. All data are presented as mean ± SEM of three independent experiments.To optimise the biophysical and physicochemical properties of the CPs, and to identify a suitable attachment point for half-life extending moieties, such as albumin binders, we sought to establish a detailed structure–activity relationship (SAR) on cluster B. We chose the most and least abundant of our cluster B lead peptides (B_3 and B_1275) as starting points for full amino acid mutation scans.24,26 We focussed on natural amino acids only to keep the option for (semi-)recombinant expression of the compounds, should large amounts be needed for further development stages. Single mutation variant peptides were synthesised in a 96 well format for each amino acid in the variable region contained between the two conserved cysteine residues C2 and C13. Methionine and cysteine were not included in the mutational scan due to potential oxidation, and asparagine was not included due to the risk of deamidation and isomerisation to iso-aspartate. Additionally, scrambled peptides were included in the panel. The binding affinities of these mutant peptides to hGIP-R ECD were then determined by single concentration BLI experiments.As shown in Fig. 3a and b, the consensus sequence of cluster B (positions 9–12, LWPF) was found to be the most intolerant to mutation, confirming this region to be critical for binding. Proline at position 11 was found to be completely intolerant to replacement by any other amino acid, suggesting that this residue is critical for maintaining conformationally restricted binding interface of CPs. While residues 9–12 are hydrophobic in both parent peptides, it is unlikely that the interaction of the peptides to hGIP-R ECD is unspecific, as none of the scrambled peptides (e.g. LWPF) were found to bind, including the transposed mutants of B_3 and B_1275, with interchanged L9 and W10. The data shows some general trends for both B_3 and B_1275. In general, positions 3 and 6 (both F in the parent peptides) favour aromatic residues, position 8 favours aliphatic residues, while residues 4, 5, and 7 are fairly tolerant to mutation. The single amino acid scan did not reveal any substitutions that led to remarkable improvements in binding, suggesting that these peptides may either already be optimised for binding through several rounds of mRNA-display selection, or that further binding improvements would only be achieved by synergistic action of multiple substituted residues. The latter could ideally be investigated by mRNA display based affinity maturation experiments in follow up studies.24,26 To gain insight into the binding mode of CPs to hGIP-R on the atomistic level we employed a two-step modelling protocol that first generates multiple conformations of the complex and then selects the final conformation based on stability (for details, see ESI). As our peptides were demonstrated to be 125I-GIP-competitive, we directly folded them inside the GIP binding site of hGIP-R using Rosetta (Fig. 3c and d; results for B_1275 shown). Briefly, each of the four crucial residues according to the SAR data (Leu9, Trp10, Pro11 or Phe12) was placed in the GIP binding site and the rest of the peptide was grown around it. The obtained conformations (24 000 in total) were further clustered and the representative poses of the four most populated clusters were submitted to molecular dynamics (MD) simulations (600 ns for each pose) to assess their stability. The model with the lowest L_RMSD (Root Mean Square Deviation of the ligand backbone during MD trajectory) was selected for further characterization (Fig. 3c). In this model, the C-terminal half of the peptide faces the GIP binding site forming most of the interactions with hGIP-R. This is in line with the obtained SAR data that the C-terminal region is in general less tolerant to mutagenesis compared to the N-terminus. Furthermore, the Trp10 residue overlaps with the position of Phe22 in GIP which is known to be crucial for its binding to the receptor (Fig. 3d). Finally, overall conformation of the peptide resembles that of a hairpin with the turn located on Phe6 and His7, which are the only two positions (together with the peptide termini, see below) tolerating mutagenesis to a proline.Open in a separate windowFig. 3Heat maps showing Kd (nM) values for binding of single-residue mutant peptides derived from B3 (a) and B1275 (b) to biotinylated hGIP-R ECD as determined by single-concentration BLI. The columns indicate the amino acid changes compared to the parent sequence (displayed at the top of the table). All peptides featured a C-terminal Cys and were tested in crude form. Peptides that were not tested or those where synthesis failed are marked as grey boxes. None of the scrambled mutant peptides showed any binding. (c) Atomistic model of the cyclic peptide B_1275 and GIP receptor complex suggested by molecular modelling. The LWPF sequence of the peptide is shown in red, while the rest of the molecule is coloured by the atom name (carbon in yellow, nitrogen in blue, oxygen in red, and sulphur in orange). The receptor is show in grey; the binding site residues (L35, W39, M67 and Y87) are highlighted in green. (d) Overlay of the crystal structure of GIP complexed with hGIP receptor (PDB 2QKH) and atomistic model of cyclic peptide B_1275 in the binding site of the GIP receptor. Note that the positions of W10 of the cyclic peptide and F22 of GIP overlap.Next, we wanted to use the accrued binding and SAR information to design peptides for in vivo studies. For these, peptides need to fulfil two crucial requirements. Firstly, they need to be stable in plasma and secondly, they need to be protected from renal clearance. The latter is crucial to the development of any peptide therapeutic as even protease resistant peptides usually exhibit plasma half lives in the range of a few minutes. Attachment of a fatty acid-based albumin binder (termed protractor) is one of the most employed strategies to extend peptide half-life and we opted for a 2xOEG-gGlu-C18 diacid albumin binder for our peptide hits.3,27–29 The binding model suggested that the addition of an albumin binding moiety is tolerated at the N- or C-terminus, as these are solvent exposed. Further evidence that replacement of the Met in position 1 with Ala is possible for both B_3 and B_1275 without loss of affinity and the ease of synthesis prompted us to design four peptides with a protractor attached to the N-terminus (B_3.1, B_5.1, B_68.1 and B_1275.1). Furthermore, we synthesized linear versions of the parent peptides in order to investigate the importance of macrocyclization on binding and stability (B_3.2, B_5.2, B_68.2 and B_1275.2). Additionally, we synthesized peptides without FLAG tag to prove that binding indeed is mediated by the macrocycle and not the tag (B_1275.3). During the purification of the latter, we realized that solubility was low (as expected from the rather hydrophobic sequence), complicating purification and analysis. Thus, we utilized our SAR understanding to design variants with improved biophysical properties by reducing the hydrophobicity and increasing the charge at positions where mutation was tolerated (especially positions 1, 4 and 7). Furthermore, we replaced the Met residues to avoid oxidation of the sulphur atom. We focussed on B_1275 as we anticipated from the sequence that this peptide is most hydrophilic, as also indicated by the shortest retention time in HPLC chromatography (Table S1). All resulting peptides containing a FLAG tag (B_1275.7 through B_1275.11) were soluble at the relevant conditions, however, only B_1275.5 showed high solubility without the tag, as measured by nephelometry at different PEG concentrations (Fig. S14). Pleasingly, addition of an albumin binder (B_1275.6) did not negatively affect solubility. Analysis of the peptides by multi-concentration BLI and radio-GIP displacement showed that indeed N-terminal modification was allowed for most peptides (Fig. 4). None of the linear variants (B_3.2, B_5.2, B_68.2, B_1275.2) showed binding, suggesting that macrocyclization is crucial for the interaction. Most mutant peptides, however, retained binding, including the triple mutant B_1275.9 (M1A, T4R, H7E; Kd at 31 nM) and quadruple mutants B_1275.7 (M1A, F3W, H5E, H7E, Kd at 101 nM) and B_1275.8 (M1S, T4R, H5A, H7Q, Kd at 184 nM), showing the power of having detailed SAR information available for multi-factorial optimisation of hits. Interestingly, in contrast to the binding data obtained from F11P mutants from high throughput mutagenesis studies (Fig. 3), B_1275.10 with a F11P substitution did not show any binding, which exemplifies the limits of high throughput SAR analysis using crude peptides, as the terminal P possibly interferes with cyclisation and might lead to multi- or polymers which interfere with the BLI assay.Finally, we tested a panel of the optimised peptides for stability in human plasma and determined in vivo pharmacokinetic parameters in rats. Most cyclic peptides (B_3.1, B_1275.3, B_1275.4, B_1275.5 and B_1275.6) showed no decrease in plasma stability over the course of 5 h, including in the presence of a FLAG tag or of a protractor (Fig. 4b). While parent B_1275 had a t1/2 of 3.5 h, the two linear peptides tested (B_3.2 and B_1275.2) exhibited low levels of degradation (t1/2ca. 2 h, see Table S2 in ESI for exact values), and the biologically active linear peptides of both GIP and GLP-1 were more rapidly degraded (t1/2ca. 35–45 min), which highlights the benefits of cyclic peptides in terms of druggability. Having established the plasma stability, we turned our attention to the in vivo half-life. We chose two protracted parent peptides (B_3.1 and B_1275.1) and the optimised B_1275 variants with FLAG tag (B_1275.4), without FLAG tag (B_1275.5) and with protractor but without FLAG tag (B_1275.6). The two non-protracted peptides exhibit very short half-lives (B_1275.4 = 3.5 min, B_1275.5 = 1.8 min), whereas the peptides carrying an albumin binder show significant plasma exposure over a long period of time (t1/2, B_3.1 = 9.8 h,B_1275.1 = 10.7 h B_1275.6 = 4.2 h), with half-lives being increased by more than 100 fold (Fig. 4d) (as a comparison, the GLP-1 analogue semaglutide, which is dosed once-weekly in humans has a t1/2 in rats of 7 h.)28  相似文献   

20.
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   

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