首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 109 毫秒
1.
The biosynthetic gene cluster of the antifungal metabolite sporothriolide 1 was identified from three producing ascomycetes: Hypomontagnella monticulosa MUCL 54604, H. spongiphila CLL 205 and H. submonticulosa DAOMC 242471. A transformation protocol was established, and genes encoding a fatty acid synthase subunit and a citrate synthase were simultaneously knocked out which led to loss of sporothriolide and sporochartine production. In vitro reactions showed that the sporochartines are derived from non-enzymatic Diels–Alder cycloaddition of 1 and trienylfuranol A 7 during the fermentation and extraction process. Heterologous expression of the spo genes in Aspergillus oryzae then led to the production of intermediates and shunts and delineation of a new fungal biosynthetic pathway originating in fatty acid biosynthesis. Finally, a hydrolase was revealed by in vitro studies likely contributing towards self-resistance of the producer organism.

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

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

2.
3.
Signal Amplification by Reversible Exchange (SABRE) is a catalytic method for improving the detection of molecules by magnetic resonance spectroscopy. It achieves this by simultaneously binding the target substrate (sub) and para-hydrogen to a metal centre. To date, sterically large substrates are relatively inaccessible to SABRE due to their weak binding leading to catalyst destabilisation. We overcome this problem here through a simple co-ligand strategy that allows the hyperpolarisation of a range of weakly binding and sterically encumbered N-heterocycles. The resulting 1H NMR signal size is increased by up to 1400 times relative to their more usual Boltzmann controlled levels at 400 MHz. Hence, a significant reduction in scan time is achieved. The SABRE catalyst in these systems takes the form [IrX(H)2(NHC)(sulfoxide)(sub)] where X = Cl, Br or I. These complexes are shown to undergo very rapid ligand exchange and lower temperatures dramatically improve the efficiency of these SABRE catalysts.

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

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

4.
Mercury porosimetry and in situ high pressure single crystal X-ray diffraction revealed the wine-rack CUK-1 MOF as a unique crystalline material capable of a fully reversible mechanical pressure-triggered structural contraction. The near-absence of hysteresis upon cycling exhibited by this robust MOF, akin to an ideal molecular spring, is associated with a constant work energy storage capacity of 40 J g−1. Molecular simulations were further deployed to uncover the free-energy landscape behind this unprecedented pressure-responsive phenomenon in the area of compliant hybrid porous materials. This discovery is of utmost importance from the perspective of instant energy storage and delivery.

Mercury porosimetry and in situ high pressure single crystal X-ray diffraction revealed the wine-rack CUK-1 MOF as a unique crystalline material capable of a fully reversible mechanical pressure-triggered structural contraction.

Reducing the world''s fossil fuel dependence is the focus of many global initiatives,1 aiming to mitigate the effects of climate change through tapping into sustainable energy resources such as solar and wind power. However, increasing reliance on these renewable energy sources has introduced difficulties due to the offset between power availability and demand peaks. Complementary technologies are necessary to alleviate intermittent supply, such as peaking power plants, demand-side energy management, or large scale energy storage.2 The latter is particularly desirable as it can decouple electricity production and consumption, however the lack of a “one size fits all” approach has led the scientific community to envisage unconventional energy storage strategies.One such avenue emerging in recent years is the storage of mechanical energy via the compression of a suitable stimuli-responsive system, either through the intrusion of a non-wetting fluid into hydrophobic porous frameworks,3 or by means of application of an external pressure on flexible materials.4The former approach, first pioneered using water intrusion in zeolites and silicas,11 has recently been extended to small pore zeolitic imidazolate frameworks.12 Unfortunately, besides requiring highly hydrophobic systems, water intrusion achieves a relatively low stored energy density,3 of around 3–25 J g−1. The second strategy takes advantage of the compliant nature of bulk materials. Energy is stored through structural deformations, manifesting as continuous or sudden volume changes under external pressure. The energy stored in flexible materials over a compression/decompression cycle can be an order of magnitude higher compared to the values achieved using fluid intrusion in rigid porous systems.13 In theory, three types of pressure-induced structural behaviour can be envisioned for such a responsive system. If the structure contraction is non-reversible, all energy is dissipated and the system is categorized as a nano-shock absorber (Fig. 1b). For structural changes that are reversible upon decompression two families of system can be distinguished, i.e. a nano-damper (Fig. 1c) or an ideal nano-spring (Fig. 1d) when the pV curves show hysteresis or fully overlap, respectively.14Open in a separate windowFig. 1(A) Schema of mechanical energy storage in compliant crystalline materials, implying a unit cell volume change between open (op) and contracted (cp) structures, and prototypical pressure-volume curves of stimuli-responsive materials under mechanical pressure for (B) nano-shock absorbers, exemplified by MIL-53(Al),5 MIL-53(Ga)-FA6 and ZIF-4(Zn),7 (C) nano-dampers e.g. MIL-53(Cr),8 MIL-47(V)9 and MIL-53(Al)-FA10 and (D) nano-springs, insofar exhibited exclusively by CUK-1 presented herein.Metal–organic frameworks (MOFs), a class of porous, crystalline materials comprised of metal vertices interconnected by organic linkers, are known to exhibit responsiveness to a variety of stimuli,15,16 including external pressure.17 Recently, several frameworks of this family of hybrid materials have been shown to act as energy storing nano-dampers or energy dissipative nano-shock absorbers, as is the case for the highly flexible MIL-53(M)5,8,10 and MIL-47(V)9 series and more recently ZIF-4(Zn)7 (see Fig. 1b and c for their related structural behaviours). In such flexible crystalline materials compression is associated with a displacive phase transition between distinct structures of differing unit cell volumes, denoted as open (op) and contracted (cp) forms15,16 and illustrated in Fig. 1a, occurring reversibly or irreversibly for a nano-damper or nano-shock absorber, respectively. The considerable stored energy associated with this transition, in the range of 30–200 J g−1 (up to 4 kJ g−1 for shock absorbers18) is highly attractive from the perspective of mechanical energy storage. However, the hysteretic compression/decompression curve characterising known nano-damper MOFs leads to a partial loss of work energy, lowering the potential storage efficiency, as well as creating issues through heat dissipation. Insofar, the search for a ideal spring-like crystalline material, capable of reversible pressure-induced structural switching without any hysteresis (Fig. 1d) has been fruitless, precluding their applicability for efficient, high density energy storage applications. Herein, a subtle combination of Hg-porosimetry, high-pressure single crystal X-ray diffraction (SC-XRD) and molecular simulations reveals the 1D-channel CUK-1 (M, M = Co, Mg)19 MOF as the first compliant hybrid porous material with a spring-back mechanical breathing behaviour.Such unique mechanically-triggered structural response implies a continuous pore contraction/expansion between op and cp forms in a narrow pressure range of 280–290 MPa, accompanied by a unit cell volume change of 20.9%. This optimal scenario paves the way towards fast energy storage/delivery system of about 40 J g−1. The channel-like CUK-1(M) composed of chains of μ3-OH/O edge and vertex sharing metal octahedra (M = Co,19 Mg20) coordinated by bidentate 2,4-pyridinedicarboxylic ligands, recently emerged as an attractive porous material owing to its promising sorption performance combined with environmentally-friendly hydrothermal synthesis and high thermal and chemical stability.20–22 Its wine-rack topology and its relatively rigid behaviour upon guest adsorption are reminiscent to that of MIL-47(V) a MOF which interestingly underwent a hysteretic, reversible structural contraction upon exerting an external pressure of 125 MPa,9 associated with a stored/delivered energy of 33 J g−1. Inspired by our previous findings on MIL-47(V), we deliberately explored the pressure-induced structural behaviour of CUK-1 in its isostructural Co and Mg forms. MOF synthesis was performed according to the protocol detailed in ESI. Phase purity was confirmed by powder XRD (Fig. S3, S4 and Table S1) while their textural features, including BET area and pore volume, were found to match previously reported data.19,20Mercury intrusion curves were recorded on the powder samples up to a maximum of 413 MPa as shown for CUK-1(Co) in Fig. 2, its Mg variant being reported in Fig. S6, ESI, together with full experimental details. A substantial amount of Hg intrudes at low pressure (<10 MPa), due to compaction of the crystals and filling of inter-particle porosity. This is followed by a sudden volume change at 281 MPa where a sharp step is observed (see inset of Fig. 2). By analogy with the conclusions previously drawn for the series of MIL-53(M)/MIL-47(V) frameworks,5,8–10 this intruded Hg volume increase is associated with a structural contraction of CUK-1(Co), as its channel size (approx. 6.6 Å) is an order of magnitude below the pore dimension where non-wetting mercury can intrude in this pressure range (at 52 Å). The extrusion curve shows a near-perfect overlap, indicating that the framework behaves as an ideal spring, with no hysteresis between the intrusion/extrusion branches.Open in a separate windowFig. 2Sequential mercury intrusion–extrusion curves on CUK-1(Co) powder, in blue line and red, respectively. Line is a guide for eye. Volume below 1 MPa corresponds to powder compaction and intercrystallite void filling. Dotted horizontal lines demarcate contraction lower and upper bounds. Inset highlights the intrusion step in a linear scale with the op/cp contraction marked with an arrow.Moreover, this behaviour is highly repeatable, as confirmed by four consecutive pressure cycles (in Fig. S5). Interestingly, the same behaviour also holds true for CUK-1(Mg) (Fig. S6), with a similar intrusion pressure of 288 MPa. Since the two metal ions show relatively similar ionic radius (Co2+: 1.50(7) Å and Mg2+: 1.41(7) Å),24 the averaged metal–oxygen distance is nearly identical in their corresponding coordination: spheres: (Co–O: 2.107(20) Å and Mg–O: 2.073(20) Å). Such analogous metal-linker bond strength is most likely at the origin of the very similar pressure-induced response of the two materials. The high transition pressure of CUK-1(Co) underpins the inability of guest adsorption to induce a breathing effect as observed previously.20,21 Indeed, the adsorption stresses encountered throughout guest insertion are simply insufficient to overcome the energetic penalty of transition.14,25 The 0.143 mL g−1 volume change associated with the observed step in the CUK-1(Co) intrusion curve corresponds to a 20.9% change in unit cell volume, lower than in the similar phenyl-based MIL-47(V) of 43%.9 However, the stored energy calculated through W = P × ΔV is 40 J g−1, 20% larger than the value reported for MIL-47(V)9 of 33 J g−1. Here, the higher pressure of CUK-1(Co) switching, 281 MPa vs. 125 MPa for MIL-47(V) balances out the ΔV term. Moreover, owing to its relatively dense framework, the volumetric energy density of CUK-1(Co) remains attractive when compared to water intrusion systems (Table S4).Considering an initial unit cell volume for the CUK-1(Co) op form of 2467 Å3 from PXRD (see ESI), the resulting cp form is estimated to exhibit a unit cell volume of 1950 Å3, based on the Hg intruded volume increase at 281 MPa. In order to directly observe the contracted form and identify the mechanism underpinning these intriguing dynamics, high pressure SC-XRD experiments were carried out in a membrane diamond anvil cell (mDAC). Individual CUK-1(Co) crystals were placed in a gasket between the polished diamonds of the mDAC, and immersed in a hydrostatic pressure transmitting medium of silicone oil AP-100, with a gold flake used to monitor inner mDAC pressure (full single crystal synthesis conditions and SC-XRD methodology available in the ESI).At ambient pressure, the indexed unit cell volume of the initial op form of CUK-1(Co) is nearly identical (2492 Å3) to that of the previously reported20 dehydrated monoclinic phase (2466.72 Å3). Upon increasing DAC pressure to around 0.3 GPa, a volume contraction to the cp phase begins, which is in line with Hg porosimetry experiments. Reflections obtained from integrated 2D diffraction images were used to solve the pressure-induced structure through a dual space recycling algorithm in an expanded P1 setting, then further refined on F2 using the SHELX suite26 (complete data treatment methodology available in the ESI). The structure maintains the same C2/c space group throughout the transition between the two forms, and as such the spring-like dynamics of the framework can be described as a continuous contraction in a narrow pressure range. Above 0.5 GPa, the cp form is attained, with further pressure application leading to a linear decrease of its unit cell volume by 4% up to 1.8 GPa (Fig. 3).Open in a separate windowFig. 3Evolution of the CUK-1(Co) unit cell volume determined through indexation of Bragg reflections as a function of applied pressure as recorded in a DAC. Unit cell parameters corresponding to each pressure point can be found in Table S5, ESI.The unit cell dimensions of the solved cp form at 0.5 GPa are provided in Table 1, alongside as-indexed pristine op form parameters with Fig. 4a illustrating the two structures. The anisotropic transition is similar in nature to that of MIL-53(M)/MIL-47(V), characterised by a compaction in the b-direction (from approx. 13 Å to 9 Å) and an elongation along the a-axis (from 18 to nearly 20 Å). The change in the c-parameter is minimal, with only a slight increase, as it lies in the plane of the highly rigid octahedrally coordinated metal chains. A lowering of the angle (from 103 to 99°) is also observed, as the 1D parallel pores are straightened via the linker-induced torsion. A table comparing specific atomic distances, angles and torsions in the two forms is available in Table S7, ESI. The unit cell volume of the identified cp phase at 0.5 GPa of 1972 Å3 is only slightly higher than the value estimated from porosimetry measurements (as 1950 Å3). We attribute this offset to the different interactions of the crystal surface with the respective pressure transmitting medium (mercury vs. silicone oil), as observed previously.5Crystallographic data of the pristine (op) and high pressure (cp) phases as determined from the CUK-1(Co) SC-XRD
Formopacpb
Crystal systemMonoclinicMonoclinic
Space group C2/c (no. 15) C2/c (no. 15)
a (Å)18.024(15)19.6026(9)
b (Å)13.089(3)9.2017(9)
c (Å)10.883(13)11.0368(3)
β (°)103.92(12)99.434(3)
Unit cell volume (Å3)2492(4)1963.8(2)
Open in a separate windowaAs indexed before pressure application.bAs solved at 0.5 GPa, Rint = 3.08, R1 = 5.53, wR2 = 16.51.Open in a separate windowFig. 4(A) Comparison of the structures of the op and cp forms of CUK-1(Co), corresponding to applied pressure of 0 GPa and 0.5 GPa, respectively, with corresponding pore limiting diameters. (B) Close-up of two adjacent linkers coordinated to the metal chains, highlighting their pivoting motion: [H] a carboxylic group hinge, with the O–O axis acting as a kneecap identical to that of MIL-53(Cr),23 and [H’] a dual carboxylic group – pyridine nitrogen hinge with the axis of rotation offset through the Co atom. (C) Concerted linker motion displayed alongside the c axis, highlighting the phenyl–phenyl distances in the two forms. Cobalt atoms/tetrahedra, oxygen, carbon and nitrogen atoms are depicted in orange, red, light grey and light blue, respectively, with hydrogen atoms omitted for clarity.A careful inspection of the op/cp structures suggests that the unique spring-like behaviour of the CUK-1 framework under pressure can be related to the position and concerted motion of its linkers upon compaction. The asymmetric linker coordination of the pyridine nitrogen to the Co–O chains results in a dual-hinged pivot, while the opposing carboxylic group adopts a single hinge configuration as in MIL-53 (H′ and H in Fig. 4b).In conjunction with alternating connectivity in both the b and c-crystallographic axes, as depicted in Fig. 4c, the corrugated channels maintain a similar separating distance between adjacent linkers throughout contraction (4.7 Å to 4.4 Å). The aromatic stacking interactions are therefore much weaker than those encountered between facing phenyl rings in the cp forms of MIL-53(M)/MIL-47(V), (separating distances lower than 4 Å) which all show irreversible or hysteretic behaviours upon pressure release.8–10 Such a conformation is expected to lower the relative stability of the cp form of CUK-1, rendering possible a complete reversibility of the cp/op transition.In order to gain further insight into the mechanical behaviour of CUK-1, we computed the internal energy of CUK-1(Mg) as a function of unit cell volume at 0 K using density functional theory (DFT) calculations. The Mg variant was chosen for this computational investigation because Co is a transition metal with a high spin state in CUK-1 resulting in an electronic structure that is much more difficult to resolve using standard DFT. The calculations were performed using the Vienna Ab initio Simulation Package (VASP)27 with the PBE-D3(BJ)28,29 functional (details in the ESI) and the results are shown in Fig. 5.Open in a separate windowFig. 5Internal energy of the CUK-1(Mg) framework as a function of its unit cell volume at 0 K computed at the DFT level.The profile reveals a single minimum located at around 2500 Å3 which can hence be identified as the op form. No local minimum is present at a lower unit cell volume, which is consistent with the experimental observation of a pressure-induced reversible contraction. However, the profile does exhibit an inflection point in the volume range of 1800–2000 Å3, which fits with the dimension of the cp form observed experimentally. As was mentioned before, dispersion interactions between stacking aromatic linkers has been put forward as the driving force for structural transitions in MOFs such as MIL-53(Al). In previous theoretical work30 it was concluded that the contribution of dispersion to the DFT energy difference between the op and cp states of MIL-53(Al) heavily depends on the applied level of theory, making it very difficult to make quantitative statements about dispersion contribution from DFT calculations at this point. However, we herein assume we can still make a qualitative comparison between two different materials (i.e. MIL-53(Al) and the present CUK-1) using the same level of theory. In this light, we found from the aforementioned theoretical work that the dispersion contribution for MIL-53(Al) at the PBE-D3(BJ) level of theory amounts to 138 J g−1 (145 kJ mol−1 per unit cell) which is responsible for stabilizing the cp state ultimately resulting in an irreversible transition. In the present case of CUK-1, our DFT calculations using the same functional showed that the contribution of the dispersion to the energy difference between the op state (chosen at 2500 Å3) and the cp state (chosen at 1900 Å3) is only 84 J g−1 (110 kJ mol−1 per unit cell). As the dispersion contribution for CUK-1 is much lower than for MIL-53(Al), we can interpret this as a lower degree of stabilization of the cp form. This, in combination with the deformation energy required for the linker staggering that was mentioned before and illustrated in Fig. 4, prevents the contracted form from being a metastable state and hence promotes the ideal spring-like behaviour.As the energy profile exhibits an inflection point at the contracted pore volume range, a volume versus pressure response derived from this profile would reveal hysteresis in the transition pressures, which is not in agreement with the ideal-spring behaviour observed experimentally. However, the DFT profile was computed at 0 K, while the experiments were performed at room temperature. Therefore, we derived a force field for CUK-1(Mg) using QuickFF31 (more details on the force field derivation can be found in the ESI) and computed the free energy at 300 K as a function of unit cell volume from molecular dynamics simulations using a protocol outlined in earlier work.14 The corresponding pressure profile (shown in Fig. S8a of the ESI) indicates a reversible transition lacking any hysteresis, in excellent agreement with the experimentally observed ideal-spring behaviour for this solid. Finally, upon decomposing the free energy into internal energy and entropy (see Fig. S8b of the ESI), we found that the entropic contributions stabilize the op phase (or equivalently destabilize the cp phase), which further enhances the ideal spring-like behaviour.  相似文献   

5.
Ellman''s reagent has caused substantial confusion and concern as a probe for thiol-mediated uptake because it is the only established inhibitor available but works neither efficiently nor reliably. Here we use fluorescent cyclic oligochalcogenides that enter cells by thiol-mediated uptake to systematically screen for more potent inhibitors, including epidithiodiketopiperazines, benzopolysulfanes, disulfide-bridged γ-turned peptides, heteroaromatic sulfones and cyclic thiosulfonates, thiosulfinates and disulfides. With nanomolar activity, the best inhibitors identified are more than 5000 times better than Ellman''s reagent. Different activities found with different reporters reveal thiol-mediated uptake as a complex multitarget process. Preliminary results on the inhibition of the cellular uptake of pseudo-lentivectors expressing SARS-CoV-2 spike protein do not exclude potential of efficient inhibitors of thiol-mediated uptake for the development of new antivirals.

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

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

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

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

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

7.
8.
We report a method for the enantioselective hydrogenation of annulated arenes using 4H-pyrido[1,2-a]pyrimidinones as substrates. The method selectively generates multiple stereocenters in adjacent rings leading to architecturally complex motifs, which resemble bioactive molecules. The mechanistic study of the stereochemical outcome revealed that the catalyst is able to overcome substrate stereocontrol providing all-cis-substituted products predominantly. In a sequential protocol, a matching interaction between catalyst and substrate stereocontrol is achieved that facilitates diastereo- and enantioselective access to trans-products.

We report a method for the enantioselective hydrogenation of annulated arenes using 4H-pyrido[1,2-a]pyrimidinones as substrates.

Chiral, saturated carbo- and heterocycles are important structural elements of secondary metabolites and active ingredients in drugs and agrochemicals.1 Compared to classical synthetic routes that utilize prefunctionalized precursors,2 enantioselective arene hydrogenation provides a more direct route to such motifs from readily accessible (hetero-) aromatic substrates.3 With current hydrogenation technologies, a variety of monocyclic arenes such as pyridines,4 furans,5 thiophenes,6 and annulated arenes such as quinolines and isoquinolines,7 indoles,8 and naphthalenes9 can be hydrogenated with high enantioselectivity. However, in all previous reports on the enantioselective hydrogenation of annulated arenes (1) the resulting products (2) contain only a single saturated ring with at least one aromatic sextet being preserved in the product, and consequently, with stereocenter(s) only being formed in one ring (Fig. 1A).10 An enantioselective hydrogenation of multiple annulated aromatic rings would enable the generation of multiple stereocenters at adjacent rings in a single step, leading to desirable, architecturally complex and natural product-like motifs. Herein, we report the first such example using N-bridged 4H-pyrido[1,2-a]pyrimidinone (pyrido-pyrimidinone) substrates (3, Fig. 1B).10ad Pyrido-pyrimidinones, and their saturated analogs, are featured in bioactive molecules (Fig. 1C). For instance, Roche''s Risdiplam (5), an inhibitor against spinal muscular atrophy, includes a pyrido-pyrimidinone unit11 and a semi-reduced tetrahydropyrido-pyrimidinone is a part of risperidone (6), a blockbuster drug against schizophrenia, while octahydropyrido-pyrimidinones (4) are structurally closely related to quinolizidine alkaloids such as lupanine (7). To date, only a small number of octahydropyrido-pyrimidinones (4) have been accessed with the majority resulting from the hetero-geneous, racemic hydrogenation of tetrahydropyrido-pyrimidinones (8). To the best of our knowledge, enantioenriched octahydropyrido-pyrimidinones have never been accessed by enantioselective catalysis.12Open in a separate windowFig. 1(A) Previously: only one ring of annulated arenes reduced by enantioselective hydrogenation. (B) This work: enantioselective hydrogenation of both rings of bicyclic aromatic pyrido-pyrimidinones (3). (C) Biologically active examples/analogues of pyrido-pyrimidinones and their hydrogenated products.We realized that an enantioselective hydrogenation of the readily accessible pyrido-pyrimidinones would in principle offer the most direct approach towards these valuable product motifs, provided that the chemoselectivity could be controlled and potential side products, such as tetrahydro- (8), hexahydropyrido-pyrimidinones 9 and 10, and the presumably unstable hemiaminal 11 resulting from amide reduction, could be avoided (Fig. 2).13 Furthermore, we rationalized that up to four different diastereomers could result from the creation of three independent stereocenters, making the simultaneous control of chemo-, diastereo-, and enantioselectivity a daunting challenge. Nevertheless, we were ideally positioned to overcome this challenge given the high reactivity, enantioselectivity, and broad tolerance of functional groups shown by the chiral ruthenium-bisNHC catalyst system developed by our group (12).14,15 Indeed, we observed that catalytic hydrogenation of model substrate 2,7-dimethylpyrido-pyrimidinone (3a) proceeded with near-complete chemoselectivity under a variety of reaction conditions (Table S1). Under the optimized conditions, the desired octahydropyrido-pyrimidinone was obtained in 92% yield as a mixture of just two from the possible four diastereomers (63 : 37 diastereomeric ratio (dr)) and with excellent 96% enantiomeric excess (ee) for the major diastereomer (Table S1, entry 4). The evaluation of other substitution patterns of the pyrido- pyrimidinone using the respective dimethyl-substituted substrates revealed that the 2,8-disubstituted product can also be obtained with good chemo-, enhanced diastereo-, and moderate enantioselectivity. However, substrates with 2,6- and 2,9-substitution patterns were less reactive. Only the eastern pyridine ring was reduced in 48% and 82% yield, respectively (Table S2). As such we decided to focus on substrates with a 2,7-substitution pattern when exploring the scope of this reaction (Fig. 3). Next, the scope was evaluated (Fig. 3). Substrates without a substituent in the C2-position (3b), and substrates containing sterically more hindering substituents such as a benzyl or an iso-propyl group (3c,d) in the C2-position all delivered the products in high yields as a mixture of just two diastereomers with moderate dr and excellent ee values for the major diastereomer. The dr values increase with the steric size of the C2-substituent. A phenyl substituent in the C2-position rendered the (eastern) pyrimidinone ring inactive, thus chemoselectively providing tetrahydropyrido-pyrimidinone 8e with moderate ee under the standard reaction conditions including the reaction time of 48 h. Contrastingly, the introduction of aryl groups in the 7-position of the (western) pyridine ring was well-tolerated. Consequently, octahydropyrido-pyrimidinones carrying aryl groups such as phenyl (4f), sterically demanding ortho-tolyl (4g), and other aryl groups bearing various synthetically useful functional groups such as NHBoc, OCF3, or Cl in the para-position (4h–j) can all be accessed in high yields and ee values. Curiously, while 7-para-fluorophenyl-substituted product 4k was obtained in high yield, the enantioselectivity was reduced in comparison to its chlorine-containing counterpart 4j. Products without a substituent in the 7-position (4l–n) were isolated in good yields as single diastereomers with good to high ee.Open in a separate windowFig. 2Chemo- and stereoselectivity challenges.Open in a separate windowFig. 3Scope. Combined isolated yields of both diastereomers are reported. The dr was determined by NMR-spectroscopy. The major diastereomer (and enantiomer) was assigned by X-ray crystallographic analysis of both diastereomers of 4a. The assignment for the other products was conducted by analogy. The ee of the minor diastereomer was measured in all cases (see ESI). aTraces of an unknown impurity could not be separated from the product. bProtection with trifluoroacetic anhydride. c80 bar H2 used. SINpEt: 1,3-bis(1-(naphthalene-1-yl)ethyl)-4,5-dihydroimidazolylidene; TFA: trifluoroacetate.Intrigued by those observations, we proceeded to study the stereochemical mechanism of the reaction. First, we determined the reaction progress of the hydrogenation of model substrate 3a over time (Fig. 4A). After a reaction time of 1.5 h, tetrahydropyrido-pyrimidinone 8a is almost the exclusive species before its relative abundance decreases with a coinciding increase of the proportion of octahydropyrido-pyrimidinone 4a. Notably, 7-methyl- and 7-phenyl-substituted intermediates 8a and 8f were isolated in good yields and with moderate enantiomeric excess (8a: 84% yield, 63% ee; 8f: 50% yield, 62% ee). The reaction is complete after 8 h, leaving octahydrogenated product 4a as the exclusive species. Hence 8a must be an intermediate in the hydrogenation of 3a to 4a. No further signals, other than those allocated to the substrate 3a and tetrahydrointermediate 8a, were observed in the enone or α-ketone regions at any of the time points. Furthermore, the C2 and C9a C–H bonds are cis-aligned in both diastereomers of the 2,7-dimethylproduct 4a (confirmed by X-ray crystallography, see ESI). A cis alignment would be enforced by a non-interrupted coordination of the catalyst during the hydrogenation of the (eastern) pyrimidinone ring. The codependence of the configurations of C2 and C9a would result in the formation of only two diastereomers in the hydrogenation of 7-substituted substrates and only a single diastereomer in the hydrogenation of 7-unsubstituted substrates, which is in agreement with the observations made during the exploration of the scope (Fig. 4B).16 Both observations suggest that enone 9a and imine 10a are not intermediates in this reaction. Finally, we determined that two separate effects cause the (relatively low) diastereomeric ratios of 7-substituted octahydropyrido-pyrimidinones. Firstly, the enantiocontrol over the initially formed C7-stereocenter is only moderate (8a: 63% ee after 1.5 h, Fig. 4B). Secondly, the ee of the intermediates 8 increases with the reaction progress of further hydrogenation towards the octahydroproducts 4 (Fig. 4C). It follows that the minor enantiomer of chiral intermediate 8 is converted into octahydro-product 4 more rapidly than the major enantiomer, indicating a mismatched interaction between substrate and catalyst control. To further test this conclusion, isolated intermediate 8a was submitted to the enantioselective hydrogenation using the opposite enantiomer of the chiral catalyst in the second hydrogenation step (Fig. 4D).16 In this case, hydrogenation proceeds with a matched interaction between catalyst and substrate control and provides the opposite diastereomer trans–cis-4a with excellent diastereo- and enantiocontrol. Notably even racemic arene hydrogenation rarely produces trans-products.18 The dr of all–cis-4a derived from the two-step hydrogenation (Fig. 4D, red equation) is higher than that of the standard one-pot protocol (75 : 25 vs 56 : 44 dr). This is a result of the increase of the ee of intermediate 8a as a function of the reaction time (the minor enantiomer is consumed faster to form 4a). The effective ee of 8a in the two-step protocol (isolated after 2 h) was higher than that of 8a in the standard protocol resulting in a higher dr of 4a.Open in a separate windowFig. 4Mechanistic studies. (A) Excerpt of the 1H NMR spectrum of the reaction mixture as a function of the reaction time. (B) Key observations that suggest a two-step hydrogenation process. (C) ee of 8a as a function of the reaction time. (D) Sequential enantioselective hydrogenation of 3a using opposite enantiomers of the chiral catalyst.17  相似文献   

9.
10.
11.
12.
Radical electrons tend to localize on individual molecules, resulting in an insulating (Mott–Hubbard) bandgap in the solid state. Herein, we report the crystal structure and intrinsic electronic properties of the first single crystal of a π-radical metal, tetrathiafulvalene-extended dicarboxylate (TED). The electrical conductivity is up to 30 000 S cm−1 at 2 K and 2300 S cm−1 at room temperature. Temperature dependence of resistivity obeys a T3 power-law above T > 100 K, indicating a new type of metal. X-ray crystallographic analysis clarifies the planar TED molecule, with a symmetric intramolecular hydrogen bond, is stacked along longitudinal (the a-axis) and transverse (the b-axis) directions. The π-orbitals are distributed to avoid strong local interactions. First-principles electronic calculations reveal the origin of the metallization giving rise to a wide bandwidth exceeding 1 eV near the Fermi level. TED demonstrates the effect of two-dimensional stacking of π-orbitals on electron delocalization, where a high carrier mobility of 31.6 cm2 V−1 s−1 (113 K) is achieved.

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

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

13.
14.
Amphidynamic motion in metal–organic frameworks (MOFs) is an intriguing emergent property, characterized by high rotational motion of the phenylene rings that are embedded within an open, rigid framework. Here, we show how the phenylene rings in the organic linkers of the water stable MOF PEPEP-PIZOF-2 exhibit multiple rotational rates as a result of the electronic structure of the linker, with and without the presence of highly interacting molecular guests. By selective 2H enrichment, we prepared isotopologues PIZOF-2d4 and PIZOF-2d8 and utilized solid-state 13C and 2H NMR to differentiate the dynamic behavior of specific phenylenes in the linker at room temperature. A difference of at least one order of magnitude was observed between the rates of rotation of the central and outer rings at room temperature, with the central phenylene ring, surrounded by ethynyl groups, undergoing ultrafast 180° jumps with frequencies higher than 10 MHz. Moreover, loading tetracyanoquinodimethane (TCNQ) within the pores produced significant changes in the MOF''s electronic structure, but very small changes were observed in the rotational rates, providing an unprecedented insight into the effects that internal dynamics have on guest diffusion. These findings would help elucidate the in-pore guest dynamics that affect transport phenomena in these highly used MOFs.

Zirconia based metal–organic framework PEPEP-PIZOF-2 exhibits the emergent property of amphidynamic motion with organic links that show multiple rotational rates related to their molecular structure and independently from presence of molecular guests.

Amphidynamic crystals are an emerging class of materials made of molecular components that exhibit fast internal motion within a rigid lattice.1,2 Metal–organic frameworks (MOFs) can be considered as intrinsically amphidynamic materials, because they are formed by the assembly of organic molecules that carry high degrees of freedom linked to inorganic clusters that form an extended solid matrix.3 This assembly allows for the organic components to behave like rotators, while the solid matrix/framework acts as a stator, with gyroscope-type motion enabled by the open architecture of the MOF with motion modulated by the molecular structure of the linker.4–6 In order to create materials with targeted dynamic properties for real-life applications, like molecular machines, it is important to determine whether the chemical environment of the linkers can produce dynamics at multiple rates and how the presence of molecular guests affect such dynamics. To do so, it is important to use MOFs that are chemically stable to water and humidity, because this robustness increases the reproducibility of the results and the applicability of the MOF. The interplay between guest diffusion, linker dynamics and the overall framework flexibility has been actively investigated in recent years.7Here, we prepared a water-stable MOF, PEPEP-PIZOF-2 (Fig. 1a), strategically labelled with deuterium atoms to probe the multiple segmental motion in the pristine and guest-loaded materials. Utilizing solid-state NMR techniques, we elucidated that this MOF exhibits bimodal rotational rates, with the central ring of the linker having free rotation above the 10 MHz limit of quantitation, and with the outer rings having slower rotation. This double-rate internal dynamics is preserved even in the presence of a very “sticky,” electron-deficient guest such as tetracyanoquinodimethane, TCNQ. Studying the molecular dynamics of this class of MOFs helps in accelerating their use as applied materials and for the fundamental studies of transfer phenomena that occur in MOFs such as mass, heat, and momentum transfer.Open in a separate windowFig. 1(a) Crystal structure of the double interweaved MOF PIZOF-2, showing each framework in separate colors. (b) Molecular structure of the PEPEP link. (c) Deuterium enriched linkers used in this study emphasizing the deuterium location in the link and the different chemical environments.Zirconia MOFs have been shown to exhibit unprecedented chemical stability, of which the family of Porous Interpenetrated Zirconia Organic Frameworks (PIZOFs) features superior stability combined with a unique molecular composition of their linkers.8,9 The linkers in PIZOF MOFs are linear and made by a combination of phenylene rings and ethynylene groups, where multiple chemical environments can be created around the phenylenes, thus altering their rotational behavior. Of the series, PEPEP-PIZOF-2 (hereafter PIZOF-2) is a high symmetry interweaved MOF (interweaved = interpenetrated with minimally displaced frameworks10,11) made with linkers that contain three phenylenes (P) and two ethynylenes (E) in an alternating form (hence PEPEP, Fig. 1b), creating two different types of chemical and crystallographic environments around the rotor moieties: the central phenylene ring is surrounded by two alkyne groups that provide a negligible electronic barrier for rotation and two outer phenylene rings surrounded by an alkyne and a carboxylate. So, we expect to observe significant differences in dynamics for each component of the linker.12 To properly observe the gyroscope-like rotation, protons were replaced with deuterons either in the inner ring (PIZOF-2d4, Fig. 1c) or in the outer rings (PIZOF-2d8, Fig. 1c). These two modes of isotopic labeling allowed the isolation of each ring to study of their dynamics by 2H NMR.Samples of the PIZOF-2 MOF containing natural and isotopically enriched PEPEP links were prepared from adapted published procedures (ESI).13 The MOFs were prepared via solvothermal crystallization of the respective linkers in DMF in the presence of ZrCl4 and proline-HCl at 120 °C for 24 h, resulting in crystalline powder samples of formula Zr6O4(OH)4[PEPEP]6, Zr6O4(OH)4[PEPEP-d4]6, and Zr6O4(OH)4[PEPEP-d8]6. Powder X-ray diffraction (PXRD) patterns of all three isotopologues exhibited sharp diffraction lines starting at 3.84° 2θ (CuKα radiation) characteristic of the cubic PIZOF-2 MOF phase (Fd3̄m space group symmetry) (Fig. 2a).14 Phase purity was assessed using Rietveld refinement of the experimental patterns using the single crystal unit cell data resulting in phase pure samples with low residuals (Fig. S1–S4).Open in a separate windowFig. 2(a) Powder X-ray diffraction of the natural and isotopically enriched PIZOF-2 MOFs demonstrating their isoreticular nature. Miller indices of the most intense peaks are indicated. (b) 13C CPMAS NMR spectra of natural and isotopically enriched PIZOF-2 MOFs.The internal structure of the MOFs was analyzed using 13C Cross-Polarization with Magic Angle Spinning (CP MAS) NMR spectroscopy, where the intensities of 13C signals varied according to the level of deuteration of the linker in each MOF (Fig. 2b). PIZOF-2 exhibits a 13C spectrum with signals at around 92 ppm, corresponding to the internal ethynyl, signals between 120 and 140 ppm corresponding to the phenylene carbons, and signals at 173 ppm that correspond to carboxylates, consistent with the expected structure. In PIZOF-2d4 the signals that correspond to the central phenylene ring are attenuated (Fig. 2b, signal 8) compared to the natural material, whereas in PIZOF-2d8, the only visible signals are those of the central ring, due to the absence of vicinal protons required for CP. Peaks associated with solvents and other reagents were not observed indicating a successfully evacuated framework, which in addition to high crystallinity and the magnetic field produces changes in the spectral line shape that can be associated with different types of motion.15 In the case of the PIZOF MOFs, the differences in the molecular substructure and porosity ensured having optimal samples for dynamic studies. Despite being double interweaved, the distances between centroids of the aromatic rings of the interpenetrating frameworks have values in the range of 6.23 Å to 8.04 Å (Fig. S7). Considering that the volume of revolution of the phenylene is ca. 6 Å, significant changes in the internal rotational dynamics caused by interpenetration were ruled out. Besides, it is expected that the phenylene rings have sufficient space to undergo fast rotational displacement, as it has been observed in other MOFs.16 To determine this, the deuterated samples were studied using solid-state quadrupolar echo 2H NMR spectroscopy. The reorientation of the C–2H bond vectors with respect to the external between outer and inner rings is expected to afford different rotational rates.The 2H NMR line shape at room temperature of PIZOF-2d8 displays signals characteristic of motions in the intermediate exchange regime. A successful fitting of the spectrum using NMRweblab17 was obtained using a model that assumed two-fold flip jumps, indicating a rotational rate at room temperature of the outer rings of krot = 2.10 MHz (Fig. 3 top). The rate of rotation of the deuterated outer rings is similar to that reported in UiO-66(Zr)18 (krot = 2.3 MHz at rt) and much larger than that of other simple MOFs like MOF-5(Zn),12MIL-47(V),19 and MIL-53(Cr)19 (krot < 0.001 MHz at rt). The rotation of the outer rings could be then regulated by the electronic conjugation of the phenylene with the carboxylates and/or affected by the interactions with the metal oxide clusters.Open in a separate windowFig. 3Experimental (blue) and calculated (orange) deuterium line shapes of PIZOF-2 at 295 K: (top) PIZOF-2d8 and (bottom) PIZOF-2d4.Conversely, in the case of PIZOF-2d4 (Fig. 3 bottom) the narrow 2H NMR spectrum is characteristic of ultrafast reorientations about the –C Created by potrace 1.16, written by Peter Selinger 2001-2019 C– axis. A fitting of the spectrum was carried out assuming fast 180° jumps and large amplitude vibrations, indicating a rate of rotation of krot > 10 MHz, the upper limit of the 2H NMR sensitivity, so at 295 K the inner rings are rotating freely. This rate correlates with the minimal electronic barrier given by the flanking alkynes as has also been observed in a Zn-pyrazolate MOF that contains the same diethynyl-phenylene-diethynyl moiety.16 To date, this is the first time a MOF exhibits multiple rotational rates of their phenylene rings, which has implications for understanding and improving guest-diffusion related phenomena such as guest storage, catalysis, and separations.As the transport of guests throughout the MOF would be affected by the interactions between the guest and the static and dynamic components of the framework, we impregnated deuterated PIZOF-2 samples with tetracyanoquinodimethane (TCNQ). Given the electron-rich nature of the linker, electron deficient TCNQ was selected because it fits into the pores and has a high propensity to form strong π–π stacking bonds, often in the form of charge transfer complexes.20 In other words, TCNQ is a very sticky molecule known to affect the electronic structure of MOFs and has been used as an additive to enhance their charge conduction properties for device applications.21,22The incorporation of TCNQ into the MOF was performed by immersing MOF powder samples in CH2Cl2 solutions for a minimum of 6 h at 295 K followed by rinsing, resulting in a loading capacity of 28.6 ± 0.2 TCNQ molecules per unit cell. At this saturated state, the white crystals changed to a green color and showed a strong EPR signal with g = 2.0025 (Fig. S8c), compared to the pristine MOF. This could be attributed to a charge transfer event that produces organic radicals which overshadows the intrinsic paramagnetism of the zirconia oxoclusters.23 We also observed a quench of the emission, with a significant change in the quantum yield from ΦF = 8.5% to ΦF < 0.1% (Fig. S8d). Fluorescence quenching was expected due to the interaction of electron deficient molecules with the conjugated oligo-phenylene-ethynylene linkers that make the MOF emissive.24The 13C CPMAS spectrum of TCNQ loaded PIZOF-2d4 not only confirmed the guest within the pores (Fig. S12), but it also revealed the changes in the chemical environment around the linkers: the appearance of a second carboxylate signal around δ = 174 ppm and a second quaternary carbon signal around δ = 128 ppm, with higher intensities with an increased loading time (Fig. 4a), attributable to the interaction of TCNQ with the outer rings of the PEPEP links, closer to the Zr cluster. Surprisingly, despite the evidence of the diffusion of TCNQ into the MOF, the solid-state 2H NMR spectrum of PIZOF-2d4 loaded with TCNQ for 6 h remained unaltered (Fig. 4b). Increasing the impregnation time to 72 h or increasing the temperature to 60 °C resulted in similar line shapes. These results suggest that the guest may have adsorbed near the outer phenylene rings of the linker. To demonstrate this, PIZOF-2d8 loaded with TCNQ for 6 h (Fig. 4c) was studied by 2H NMR. Interestingly, the fitting of the 2H line shape indicated slightly faster rotational rates compared to pristine PIZOF-2d8, changing from krot = 2.1 MHz to krot = 3.3 MHz. Only rising the impregnation temperature to 60 °C for 24 h allowed faster adsorption equilibration, decreasing the rotational rate to krot = 1.2 MHz. This indicates that the diffusion of TCNQ is slow and may require longer equilibration times at higher temperature to reach an equilibrium. Furthermore, considering the changes in the chemical shift of the carboxylate peak observed by 13C NMR CP MAS upon the diffusion of the guest (Fig. 4a, pink mark), as well as the minor changes in the rotational dynamics of the aromatic rings, we postulate that the TCNQ is located closer to the metal cluster, which agrees well with previously observed guest-loaded Zr-based MOFs.25,26Open in a separate windowFig. 4(a) 13C CPMAS of PIZOF-2d4 at different TCNQ loading times. (b) Experimental and simulated 2H NMR spectra of PIZOF-2d4 revealing that the signal from the central phenylene remains unaffected. (c) Experimental and simulated 2H NMR of PIZOF-2d8 under different TCNQ loading conditions.This work highlights that our approach can tackle one of the challenges in guest-loaded MOFs, which is the understanding of the interactions between the guest and the framework, a problem often exacerbated by the difficulty of acquiring high-quality single crystals. Furthermore, even after obtaining suitable crystals, X-ray diffraction studies provide only averaged space and time information. Conversely, solid-state NMR, as it is time-resolved, is ideal to analyze guest loaded MOFs in bulk samples, providing kinetic information such as transient π-interaction sites,27 gas-absorption diffusional rates,28 internal rotational dynamics,6 and other kinetic details.29,30  相似文献   

15.
In situ single-crystal diffraction and spectroscopic techniques have been used to study a previously unreported Cu-framework bis[1-(4-pyridyl)butane-1,3-dione]copper(ii) (CuPyr-I). CuPyr-I was found to exhibit high-pressure and low-temperature phase transitions, piezochromism, negative linear compressibility, and a pressure induced Jahn–Teller switch, where the switching pressure was hydrostatic media dependent.

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

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

16.
Indolmycin is an antibiotic from Streptomyces griseus ATCC 12648 with activity against Helicobacter pylori, Plasmodium falciparum, and methicillin-resistant Staphylococcus aureus. Here we describe the use of the indolmycin biosynthetic genes in E. coli to make indolmycenic acid, a chiral intermediate in indolmycin biosynthesis, which can then be converted to indolmycin through a three-step synthesis. To expand indolmycin structural diversity, we introduce a promiscuous tryptophanyl-tRNA synthetase gene (trpS) into our E. coli production system and feed halogenated indoles to generate the corresponding indolmycenic acids, ultimately allowing us to access indolmycin derivatives through synthesis. Bioactivity testing against methicillin-resistant Staphylococcus aureus showed modest antibiotic activity for 5-, 6-, and 7-fluoro-indolmycin.

A semi-synthetic system for producing indolmycin, an antibiotic, was developed and used to make indole-substituted, halogenated derivatives of indolmycin, some with modest bioactivity against methicillin-resistant Staphylococcus aureus.

Antibiotic-resistant bacteria pose a great threat to human health,1–4 and the rates of new antibiotic discoveries and clinical approvals have been in a steep decline since the 1980s.1 Without the discovery and development of new antibiotics, drug-resistant pathogens, such as methicillin-resistant Staphylococcus aureus (MRSA), will become increasingly prevalent.2–4 One strategy to increase antibiotic development has been to “rediscover” known, but underdeveloped, antibiotics.1 One such example is indolmycin, which was originally discovered in 1960 from Streptomyces griseus ATCC 126485 but was not originally developed for clinical use because of its narrow spectrum of activity6–10 and its interference with tryptophan catabolism in the liver.9,10 However, reignited interest in this old antibiotic led to the discovery of its activity against Helicobacter pylori,6Plasmodium falciparum,11 and MRSA.12 For MRSA, indolmycin was found to be active against mupirocin- and fuscidic acid-resistant MRSA strains, with strains resistant to indolmycin emerging infrequently and with reduced fitness compared to sensitive strains.12 In addition, indolmycin has been shown to have minimal activity against common members of the human microbiota, suggesting that its narrow spectrum of activity is an asset.6 The first indole-substituted derivatives, 5-hydroxy and 5-methoxyindolmycin, were made by precursor-directed feeding of the indolmycin producer, Streptomyces griseus ATCC 12648, and they showed modest improvements in bioactivity against S. aureus and Escherichia coli.13 Two practical synthetic routes to indolmycin and some indole-substituted derivatives have been reported more recently,14,15 which enabled access to a small variety of indole-substituted derivatives. Additionally, a previous patent has described synthetic methods to produce a variety of derivatives; however, these methods do not appear to offer stereochemical control, and some require tailoring steps specific to each analog.16 Therefore, further development of indolmycin would benefit from a simpler diversification method that could be applied to produce a wider variety of analogs with stereochemical control.Inspired by early biosynthetic studies,17,18 our group previously identified the indolmycin gene cluster and elucidated the biosynthetic pathway, demonstrating that indolmycin (1) is assembled from tryptophan, arginine and S-adenosylmethionine (SAM) in a three-part process (Fig. 1).19 In the first part, l-arginine is oxidized by Ind4 in an oxygen- and PLP-dependent reaction to 4,5-dehydro-2-iminoarginine, which is then enantioselectively reduced by imine reductase Ind5 and its chaperone Ind6 to 4,5-dehydro-d-arginine. In the second part, tryptophan (2) is deaminated by PLP-dependent transaminases, giving indole pyruvate (3). Compound 3 is then methylated by SAM-dependent C-methyltransferase Ind1 to 3-methyl-indolepyruvate (4) which is reduced by NADH-dependent ketone reductase Ind2 to form indolmycenic acid (5). Then, in the third part, 4,5-dehydro-d-arginine and 5 are coupled in an ATP-dependent fashion by Ind3 and Ind6, resulting in an oxazolinone-cyclized molecule, N-desmethyl-indolmycin, which is finally N-methylated by Ind7, a SAM-dependent N-methyltransferase, to form 1.Open in a separate windowFig. 1Indolmycin biosynthesis from Streptomyces griseus ATCC 12648. (a) Indolmycin biosynthetic gene cluster. (b) Indolmycin biosynthetic pathway from Streptomyces griseus ATCC 12648. (c) Semi-synthetic scheme towards indolmycin and derivatives using indolmycin biosynthetic genes. The dashed arrow indicates a predicted side-product based on LC-MS analysis.Armed with the elucidated biosynthetic pathway for 1, we set out to create an in vivo system to make 1 in E. coli. We first cloned all necessary genes into four plasmids and co-expressed the genes in E. coli (Fig. S1), a strain which we named E. coli I1234670P5 (Table S1). We found that the genes needed to produce indolmycin in E. coli were ind1, ind2, ind3, ind4, ind6, ind7, ind0 and pel5, a homologous gene of ind5 from Paenibacillus elgii B69 showing better production of active protein in E. coli.20,21 We also relied on the activity of endogenous E. coli aminotransferases to catalyze the initial tryptophan deamination step. However, only a small amount of 1 was produced (∼170 μg L−1 of bacterial culture) and the yield could not be improved despite our best efforts (Fig. 2a). However, we found that this construct produced substantial amounts of 5 ([M + H]+ = 220 m/z) at 40–50 mg L−1 of culture, along with a shunt product, C-desmethyl-indolmycenic acid (6; [M + H]+ = 206 m/z).Open in a separate windowFig. 2Biosynthetic production of 5 and semi-synthetic production of 1. (a) Extracted ion chromatograms show production of 5 with minimal production of 1 from E. coli I1234670P5. (b) Synthetic scheme to 1 from 5, adapted from literature methods.14,15 (c) Total ion chromatogram of compound 1 isolated after semi-synthesis and final purification by semi-preparative HPLC. Compounds are indicated with coloured boxes and numbered.Compound 5 itself has been a focus of total synthetic efforts toward 1, as it is the key chiral precursor.14,15,22–26 Since production of 5 was much higher than that of 1 from E. coli I1234670P5, we pursued a semi-synthetic method of obtaining 1 using our biosynthetic platform to access 5, combined with a three-step chemical transformation (Fig. 2b). We attempted to remove extraneous genes from our biosynthetic platform by only including ind0, ind1, and ind2, but the changes resulted in reduced amounts of 5 (Fig. S2). At this time, it is unclear which of the other genes may be contributing to the production of compound 5. Therefore, we employed the full eight-gene construct toward synthesis of 5. We then adapted the three-step synthesis to make indolmycin,14,15 in which purified 5 was esterified to make the ethyl ester (7; [M + H]+ = 248 m/z; Fig. S3a), cyclized to give N-desmethyl-indolmycin (8; [M + H]+ = 244 m/z; Fig. S3b and S4a), and methylated at the exocyclic nitrogen to give 1 ([M + H]+ = 258 m/z; Fig. 2, S4b and Table S4).Then, we wanted to make derivatives of 1 from indole derivatives, which are more widely accessible than derivatives of 2. The tryptophan synthase (TrpS) from Salmonella enterica has been previously shown to couple a wide variety of indole derivatives to l-serine to generate derivatives of 2.27 We were able to replace pel5 with trpS in our biosynthetic platform without a reduction in the amount of 5 produced (Fig. S1b), and we named the resulting strain E. coli I1234670TS. When we fed 5-fluoroindole to E. coli I1234670TS, we observed increased production of fluorinated metabolites, 5-fluoro-indolmycenic acid (5F-5; [M + H]+ = 238 m/z) and 5-fluoro-C-desmethyl-indolmycenic acid (5F-6; [M + H]+ = 224 m/z) (Fig. 3a). We then optimized the feeding conditions, finding that 5F-5 amounts were optimal when we fed E. coli I1234670TS with 0.5 mM 5-fluoroindole per day over two days (Fig. S5).Open in a separate windowFig. 3Addition of the trpS gene to the biosynthetic platform allows incorporation of substituted indoles into 5. (a) LC-MS analysis of strains with and without trpS (E. coli I1234670TS and E. coli I1234670P5, respectively) when fed 5-fluoroindole. Chemical structures of compounds with the corresponding extracted ion chromatogram are shown to the right of the traces. (b) Indole derivatives tested. Dark blue shows indoles that were incorporated into an analog of 5 (>48% of underivatized 5 by LC-MS analysis; Table S5) and further purified; medium blue indicates indoles that were incorporated into an analog of 5 at lower levels (6–22% of underivatized 5 by LC-MS analysis; Table S5) but were not further verified by purification; and light blue indicates indoles that did not show detectable incorporation into 5.To determine the scope of indole derivatives accepted by our biosynthetic platform, we fed a variety of indoles to E. coli I1234670TS and monitored the production of 5 and its derivatives by LC-MS. Out of the indoles tested, we found that fluorinated and chlorinated indoles substituted at the 5-, 6- and 7-positions were the best accepted by the biosynthetic platform (Fig. 3b, S6 and S7). We predict that lower acceptance of indoles substituted at the 4-position may be due to steric hindrance, as 4-fluoroindole was moderately accepted, while 4-chloroindole was not observed at all. Although we observed LC-MS peaks consistent with conversion of some of the azaindoles and hydroxyindoles into 5 derivatives, further work is required to confirm, optimize and scale up the purification of these compounds (Fig. S7). Cultures producing derivatives of 5, substituted at 5-, 6- and 7-positions, were further scaled up for purification of the 5-derivatives and downstream synthesis of 1-derivatives (Fig. S8–S10 and Table S4). Each purified derivative of 5 and 1 was characterized by HR-MS and NMR (Table S3 and ESI Methods). Overall, cultures fed with the fluorinated indoles produced a higher amount of 5-derivatives than the cultures fed with the chlorinated indoles. 1 and its derivatives were tested against MRSA (Fig. S11). While the fluorinated derivatives showed bioactivity, the chlorinated derivatives of 1 did not show bioactivity at the maximum amount tested in the disk diffusion assay (30 μg). We determined MIC50 values for each fluorinated compound (Table 1). The MIC50 values demonstrate that 1 is a more potent inhibitor of MRSA than its derivatives, while 6F-1 showed the most potent inhibition of MRSA compared to any of the derivatives, followed by 7F-1 and 5F-1. The lack of bioactivity of the chlorinated compounds may be due to the bulky chlorinated substituent hindering the compounds'' abilities to bind to the tryptophanyl-tRNA synthetase (TrpRS) target, which is supported by docking studies of the analogs into a bacterial TrpRS structure (Fig. S12).MIC50 values determined for 1 and its derivatives against MRSA. Values represent the average of three replicates ± the standard deviation. For 1, 5F-1, 6F-1 and 7F-1, concentrations ranging from 50 μg mL−1 to 0.128 ng mL−1 were tested, and for 5Cl-1, 6Cl-1 and 7Cl-1, concentrations ranging from 200 μg mL−1 to 0.512 ng mL−1 were tested. The reported MIC50 value for indolmycin is 0.5 μg mL−1 (range: 0.125–2 μg mL−1) against MRSA12
MIC50 (μg mL−1)
1 1.21 ± 0.04
5F-1 32.5 ± 19.6
6F-1 6.49 ± 0.03
7F-1 16.7 ± 4.2
5Cl-1 >200
6Cl-1 >200
7Cl-1 >200
Open in a separate windowHere we have shown that 5-, 6- and 7-fluorinated and chlorinated indole derivatives are accepted as substrates by S. enterica TrpS, E. coli tryptophan aminotransferase, Ind1 and Ind2. The fluorinated indoles, particularly 5-fluoroindole, showed greater incorporation into 5 than the chlorinated indoles, perhaps because the larger chlorine atom may have steric hindrance in some of the enzyme active sites, reducing the efficiency of turnover. This result is consistent with previous studies on TrpS from S. enterica, which shows the lowest acceptance for 4-chloro, 5-chloro and 5-bromo indoles.27 Other tested indoles, 5- and 6-hydroxyindoles and 7-azaindole, were only minimally converted to derivatives of 5 by this semi-synthetic platform, and 4-hydroxyindole and 4-, 5-, and 6-azaindole were not converted at all. Azatryptophans have been previously made using TrpS but required longer reaction times due to reduced nucleophilicity,28–30 which may explain our results. For 4-hydroxyindole, previous studies of hydroxytryptophan production by S. enterica TrpS are limited;29 however, E. coli TrpS was shown to accept 4-hydroxyindole during the production of psilocybin.31 It is unclear at this time why the 4-hydroxylated derivative of 5 was not observed. The substrate acceptance and 5-derivative production could be further improved with optimized reaction and purification conditions and engineering of strains and enzymes. For instance, TrpS has already been the target of many enzyme engineering studies focused on improving the substrate scope.32–38Previous work by Demain and co-workers demonstrated successful production of 5-hydroxy and 5-methoxy derivatives of 1 from S. griseus fed with the corresponding derivatives of indole or 2; however, 6-substituted derivatives were not converted to derivatives of 1.13 By contrast, our system incorporated halogenated substituted indoles at the 6-position, meaning that the semi-synthetic method overcomes one limitation of the purely biosynthetic method. Development of this semi-synthetic method has also helped to identify some key hurdles to creating a purely biosynthetic method. These hurdles, such as accumulation of 5, an apparent role of proteins in Parts I or III (Fig. 1) in the formation of 5, and substrate scope bottlenecks, could be addressed in future attempts to produce 1 from E. coli. Furthermore, the biosynthetic portion of our system also overcomes challenges associated with synthetic methods. Synthetic studies were typically associated with making 5 through methods such as using a chiral auxiliary,25 epoxide ring-opening,14,24 lipase-assisted enantioselective acetylation,15 or a combination of these methods,26 and few attempts to synthesize halogenated derivatives have been reported, as the methods employed are likely to be incompatible with halogenated indole precursors. A patent has described methods to produce several derivatives of 1, including 4Cl-1 and 6F-1, but stereocenters were not controlled, to the best of our knowledge, and some analogs utilized additional tailoring steps or entirely unique routes.16 Therefore, these patented methods, individually, are not robust enough to enable production of a wide variety of derivatives, making a broad search for potentially bioactive candidates challenging. Our semi-synthetic method of obtaining 1 and its derivatives offers a simple and halogen-compatible alternative to purely synthetic methods that could be used to expand and diversify the accessible indolmycin molecules for broad bioactivity searches. Broad bioactivity searches could be followed by optimization of this platform, or creation of another, to produce enough amounts of the desirable compounds to continue work beyond the initial bioactivity testing. We note that this type of semi-synthetic method could also be applied to the diversification of other tryptophan-derived molecules. Moreover, the fluorinated derivatives of 1 showed bioactivity against MRSA, demonstrating that these compounds may serve as useful molecules for development of new antibiotics. Overall, this work lays a foundation for making derivatives of 1, while demonstrating the value of combining synthetic and biosynthetic methods to make and diversify natural product-derived compounds.  相似文献   

17.
18.
19.
20.
A dual catalytic chemo-selective cross-coupling reaction of two enals is developed. One enal (without α-substitution) is activated by an NHC catalyst to form an acylazolium enolate intermediate that undergoes Michael-type addition to another enal molecule bearing an alkynyl substituent. Mechanistic studies indicate that non-covalent interactions between the alkynyl enal and the NHC·HX catalyst play important roles in substrate activation and enantioselectivity control. Many of the possible side reactions are not observed. Our reaction provides highly chemo- and diastereo-selective access to chiral lactones containing functionalizable 1,3-enyn units with excellent enantioselectivities (95 to >99% ee).

An NHC-catalyzed dual activation of two different enals is disclosed with both covalent and non-covalent activation pathways involved.

The development of chemo-selective reactions of two or more substrates bearing similar functional groups remains a classic challenge in organic synthesis.1 Enals (α,β-unsaturated aldehydes) are common building blocks that offer multiple useful modes of reactions. For instance, enals are readily used as Michael acceptors in many reactions including organic catalytic reactions mediated by amines.2 In the area of N-heterocyclic carbene (NHC) organocatalysis,3 enals are used as precursors of several NHC-bound intermediates, including Breslow acyl anion intermediates,4 homoenolate intermediates,5 enolate intermediates,6 and acylazolium intermediates.7 Somewhat surprisingly, on the other hand, there is little success in using enals as Michael acceptors to react with any of these NHC-bound intermediates.8 Elegant studies in this direction are from Scheidt, in which they showed that in the presence of an NHC catalyst, a homo coupling reaction of enals (with one enal molecule as the Michael acceptor) occurred effectively (Fig. 1a, top side).8a,c Berkessel reported an intramolecular reaction of two enal moieties (in one molecule) to form a bicyclic lactone adduct in the presence of an achiral NHC catalyst (Fig. 1a, bottom side).8b To the best of our knowledge, the intermolecular Michael addition reaction of two different enal substrates mediated by NHC catalysts has not been reported.9 Possible reasons for the difficulties of enals to behave as effective Michael acceptors likely include: (a) the relatively low electrophilicity of the α,β-unsaturated bonds of enals under the typical NHC catalytic conditions and (b) the presence of competing reactions involving both the alkene and aldehyde moieties of enals.Open in a separate windowFig. 1NHC-catalyzed reactions (a) with enals as Michael acceptors, (b) via cross intermolecular reactions of two enals, and (c) bio-active molecules bearing alkyne units.Here we disclose the first cross intermolecular reaction of two enals catalyzed by NHC catalysts (Fig. 1b). We envisioned that installation of an alkynyl substituent at the α-position of an enal can likely promote its reactivity as a Michael acceptor.10 The presence of an α-substituent can interrupt π-conjugations and thus minimize its reactivity via the corresponding enal-derived enolate/homoenolate intermediate formed with NHC, as shown by Bode, Glorius and others.6b,11 In addition, the alkynyl substituent can promote hydrogen-bonding interactions to increase the electrophilicity of the enal to react as a Michael acceptor, as observed in Jørgensen''s amine-catalyzed reactions.12 In our present study, a non-linear effect was observed regarding enantiomeric excesses of the NHC catalyst and the catalytic reaction product. The reaction enantioselectivity was also found to be sensitive to solvents and bases. These results suggested that the NHC and its azolium salt pre-catalyst (NHC·HX) played dual roles in our reaction: one is to activate the α-unsubstituted enal via the formation of the NHC-bound enolate intermediate,6 the other is to activate the α-alkynyl substituted enal via the acidic proton of the chiral NHC·HX (Fig. 1b, intermediate I & transition state TS-I).13 With respect to applications, carbon–carbon triple bonds are found in a good number of bioactive molecules such as cleviolide, (+)-prelaureatin, and oxamflatin (Fig. 1c).14 We demonstrated that our products containing these alkynyl units could be readily transformed into a diverse set of molecules.Cinnamaldehyde 1a and α-alkynyl enal 2a were chosen as the model substrates to search for suitable cross coupling reaction conditions (Table 1). The reactions were first carried out with Et3N as the base and THF as the solvent. When aminoindanol derived azoium salt A15 was used as the NHC pre-catalyst, the desired formal [4 + 2] product (3a) was obtained in a very encouraging yield (52%) with excellent ee and dr values (entry 1). The reactions appeared to be very sensitive to the structure of the NHC pre-catalysts, as similar azolium salts with N-phenyl or N–C6F5 substituents (B16 and C17) were completely ineffective, leading to no product formation (entries 2 & 3). Additional studies on the NHC pre-catalysts finally revealed that introduction of a Br substituent in the indane phenyl ring of the catalyst (D)18 led to 3a in 85% yield with 99% ee as nearly a single diastereomer (entry 4). Replacing Et3N with DIEA led to similar results (entry 5). Very interestingly, when the bases were replaced with DABCO or K3PO4, a significant drop in the enantioselectivity was observed (entries 6 & 7; see the ESI for more details). Changing the solvent from THF to CHCl3 or EtOAc has moderate effects on reaction yields (entries 8 & 9).Optimization of reaction conditionsa
EntryNHCBaseSolventYieldb (%)Eec (%)Drd
1 A Et3NTHF5299>20 : 1
2 B Et3NTHF0
3 C Et3NTHF0
4 D Et3N THF 85 99 >20 : 1
5 D DIEATHF8398>20 : 1
6 D DABCOTHF7267>20 : 1
7 D K3PO4THF8079>20 : 1
8 D Et3NCHCl36497>20 : 1
9 D Et3NEtOAc6899>20 : 1
Open in a separate windowaUnless otherwise specified, the reactions were carried using 1a (0.15 mmol), 2a (0.1 mmol), NHC (0.02 mmol), base (0.05 mmol) and solvent (1.0 mL) at rt for 24 h.bIsolated yield of 3a.cThe ee values were determined via HPLC on a chiral stationary phase.dDr values were determined via1H NMR of the crude reaction mixture.Our reactions are highly chemo-selective. Under all these conditions (Table 1), several possible side products were not formed. For example, possible adducts with enal 2a as the enolate precursor (to form 3a′ or 3a′′′) were not observed. This is not a complete surprise as α-substituted enals are unreactive azolium enolate intermediate precursors under NHC catalysis.11 Our results showed that mixing of enal 2a with highly reactive electrophiles (such as alkylidene diketone; see the ESI for more details) did not lead to any formal [2 + 4] addition product. Interestingly, the simple enal 1a did not behave as a Michael acceptor under our conditions, as homo-coupling adduct 3a′′ was not observed. In Scheidt''s elegant study, the introduction of a Lewis acid additive is necessary to activate one molecule of the enal to react as a Michael acceptor.8aOur further control experiments showed that when the α-alkynl substituent of 2a was replaced with an alkyl (e.g., Fig. 2, 2a1), vinyl (2a2), phenyl (2a3) or cyano (2a4) unit, the corresponding cross [2 + 4] reactions were not observed, with most of the starting materials remaining unchanged (for more details, see the ESI). It is clear that the alkynl unit present in enal 2a played more important roles than simply blocking the enal α-carbon to interrupt the π-conjugations. Although attempts to identify key intermediates (and possible non-covalent interactions) between the NHC catalysts and the two enals did not lead to conclusive mechanistic pictures, our experiments did show strong non-linear effects with respect to the optical purities of the NHC pre-catalyst and the reaction product (Fig. 3, see the ESI for more details).Open in a separate windowFig. 2Unsuccessful α-substituted enal substrates for the NHC catalytic chemo-selective cross [2 + 4] reactions.Open in a separate windowFig. 3Nonlinear effects with respect to the product ee and the catalyst ee values using different bases: (a) Et3N and (b) DABCO.Specifically, the reaction of 1a and 2a was studied by varying the enantiomeric purities of the NHC pre-catalyst D under the optimized reaction conditions as indicated in Table 1, entry 4 (Fig. 3). The ee values of the products and the ee values of the catalysts showed an obvious negative nonlinear effect (Fig. 3a). This nonlinear effect suggests that at least two catalysts are involved in the enantio-differentiating step of our reaction.19 It appears both of the enals (1a and 2a) are activated by NHC and/or its salt (NHC·HX) in our formal [2 + 4] reaction. It is well established that cinnamaldehyde (1a) can be activated by NHC to form an acylazolium enolate intermediate.6 We therefore propose that the other enal (2a) bearing an alkynyl unit is activated by the acidic proton from NHC·HX via non-covalent interactions. These non-covalent interactions between 2a and NHC·HX could be further supported by the “linear-effect” shown by the ee values of the products and the catalysts when using DABCO as the base (Fig. 3b). In this case, only one catalyst was involved in the enantio-differentiating step of our reaction, since the non-covalent H-bonding interactions between 2a and NHC·HX could be broken by a stronger base (e.g., DABCO, K3PO4, see the ESI for details) existing in the catalytic system. Similar activation of the α-alkynyl enal by a proton was proposed in Jørgensen''s amine-catalyzed reaction.11 In the field of NHC related catalysis, the use of NHC·HX as a H-bond donating catalyst has been demonstrated by Huang, Scheidt, Guin, and others.13The non-covalent interactions between the NHC pre-catalyst D and the alkynyl enal 2a can also be supported by 1H NMR analysis (Fig. 4). In the presence of the weak base Et3N, the acidic proton of the NHC pre-catalyst D shows an obvious change in the chemical shift after mixing with the alkynyl enal 2a (Fig. 4, a vs. b). Meanwhile, the chemical shift of the aldehyde proton of the substrate 2a is not changed in the same reaction system (a vs. c). These results support the existence of a non-covalent interaction between the NHC pre-catalyst D and the alkynyl enal 2a in our NHC organocatalytic reaction system (for more details, see the ESI).Open in a separate windowFig. 4Chemical shift of the acidic proton of the NHC pre-catalyst D under various conditions.We then examined the substrate scope using different enals (1) to react with 2a under the optimized reaction conditions indicated in Table 1, entry 4 (Scheme 1). Substituents could be installed at each position of the phenyl ring of the cinnamaldehyde 1a, with all the products afforded in moderate to excellent yields with excellent chemo-, enantio- and diastereoselectivities (3b to 3p). The β-phenyl rings of the enal substrates (1) could also be switched to a naphthyl group or heteroaromatic groups. The corresponding products were afforded in excellent enantioselectivities, although the yields or dr values dropped in these cases (3q to 3s). To our delight, aliphatic enals could also be used as the enolate precursors for this NHC catalyzed chemoselective reaction, with the desired products afforded in moderate yields with excellent dr and ee values (3t & 3u).Open in a separate windowScheme 1Scope of enals 1. aReaction conditions as stated in Table 1, entry 4. Yields are isolated yields after purification by column chromatography. Er values were determined via HPLC on a chiral stationary phase. bThe reaction was carried out on a 1.0 mmol scale based on 2a.The scope of the α-alkynyl enal substrates (2) was also examined (Scheme 2). Electron-donating substituents could be well tolerated on the β-phenyl rings of the α-alkynyl enals, with the desired products afforded in good yields with excellent ee values as single diastereomers (4a & 4b). The yields of the [2 + 4] products decreased when installing electron-withdrawing groups at any position of the β-phenyl rings, although the enantioselectivities were not affected (4c to 4f). The β-phenyl rings of the α-alkynyl enal substrates (2) could also be replaced with various heteroaromatic groups without obvious reduction in the product yields or stereoselectivities (4g & 4h). Substituents were also well tolerated on the phenyl rings attached to the alkynyl units of the enal substrates 2, with all the corresponding products afforded in moderate to good yields with excellent optical purities as single diastereomers (4i to 4p). Enal substrates 2 bearing heteroaromatic, aliphatic or terminal α-alkynyl groups also worked well in this reaction and gave the target products in moderate to good yields with excellent enantio- and diastereoselectivities (4q to 4w).Open in a separate windowScheme 2Scope of α-alkynyl enals 2. aReaction conditions as stated in Table 1, entry 4. Yields are isolated yields after purification by column chromatography. Er values were determined via HPLC on a chiral stationary phase. bThe reaction was carried out on a 6.4 mmol scale based on 2v (1.0 g).As a technical note, this chemo-selective reaction of α,β-unsaturated enals could be carried out on a large scale without reduction of the product ee or dr values, although the yields of the final products slightly dropped (e.g., Scheme 1, 3a & Scheme 2, 4v).Having examined the reaction scope with both of the enal reactants, we next seek to get additional insights into the reaction mechanism. Hammett studies20 were carried out using alkynyl enal substrates 2 bearing various p-substituents on the phenyl groups of the alkynyl units (Fig. 5). Alkynyl enal substrates 2 bearing 4-F (2i), 4-Cl (2j), 4-CF3 (2k), 4-CH3 (2l), and 4-OCH3 (2m) groups were chosen as the target substrates to evaluate their relative reaction rates compared with the alkynyl enal 2a. Kinetic studies showed that the substrates 2 bearing electron-withdrawing groups reacted faster than those bearing electron-donating groups (Fig. 5a). The Hammett plot of the relative reaction rates of the substrates 2i to 2m gave a positive slope (ρ = 1.0128). Therefore, a negatively charged transition state should be built up in the rate determining step of this [2 + 4] cycloaddition process. This is in accordance with the non-covalent H-bonding interactions that we have proposed to exist between the acidic proton of the NHC-precatalyst D and the alkynyl unit of the enal substrate 2 (Fig. 1b, TS-I, see the ESI for more details).Open in a separate windowFig. 5(a) Kinetic data and (b) Hammet plot for the competitive [2 + 4] cycloaddition reactions based on the σ values.Additionally, substrates 2x and 2y bearing steric bulky substituted phenyl groups on the alkynyl units were further examined for this NHC dual catalytic [2 + 4] cycloaddition reaction (Fig. 6). It is not surprising that the corresponding reaction products 4x and 4y were only afforded in poor yields with moderate ee values. Because the alkynly groups of the substrates 2x and 2y were shielded by the bulky mesityl and 2,6-diisopropylphenyl groups, the H-bonding interactions between the NHC pre-catalyst D and the alkynyl groups cannot be efficiently formed in these cases.Open in a separate windowFig. 6Reactions with enals 2 bearing bulky alkynyl substituents.The chiral alkynyl pyranone products obtained from this methodology are rich in functionalities for further synthetic transformations (Fig. 7). For instance, the alkynyl group in 3a could react with the adjacent phenyl group under the catalysis of Cu(OTf)2 to give tricyclic product 5 in a good yield without reduction of the optical purity.21 The terminal alkylnyl group in 4v could participate in various addition reactions and afford a variety of multi-functionalized alkene products in moderate to excellent yields with excellent ee values as single diastereomers (e.g., 6, 7, 8, 10).22 A click reaction between the alkynyl group in 4v and benzyl azide led to the formation of the chiral triazole product 9 in almost quantitative yield with excellent optical purity as a single diastereomer.22d The ethynyl group in 4v could be selectively reduced to an ethyl group with a Pd/C and CaCO3 catalyst in a hydrogen atmosphere (11). Pyranone 4w bearing a 2-trimethylsilylethynyl group could be coupled with 2-iodoaniline to give the indole product 12 in a moderate yield and diastereoselectivity with an excellent ee value.23Open in a separate windowFig. 7Synthetic transformations of the chiral pyranone products. aCuBr2, CH3CN, r.t.,1 h; bTogni reagent, TMSCN, Cu(OAc)2, terpyridine, CH3CN, 70 °C, 5 h; cTosNHNH2, FeCl3, TBHP, CH3CN, 80 °C, 8 h; dBnN3, sodium l-ascorbate, DCM/H2O (v/v = 1/1), r.t., 12 h; eNaI, TMSCl, H2O, CH3CN, r.t., 4 h; fPd/C, CaCO3, H2 (balloon), EtOH, r.t., 2 h.  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