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 15 毫秒
1.
Since the end of 2010, more than 20 synthetic cannabimimetics have been identified in ‘Spice’ products, demonstrating the enormous dynamic in this field. In an effort to cope with the problem, many countries have already undertaken legal measures by putting some of these compounds under control. Nevertheless, once a number of compounds were scheduled, they were soon replaced by other synthetic cannabinoids. In this article, we report the identification of a new – and due to its substitution pattern rather uncommon – cannabimimetic found in several ‘herbal incense’ products. The GC–EI mass spectrum first led to misidentification as the alpha‐methyl‐derivative of JWH‐250. However, since both substances show different retention indices, thin‐layer chromatography was used to isolate the unknown compound. After application of nuclear magnetic resonance spectroscopy, high‐resolution MS and GC–MS/MS techniques, the compound was identified as 3‐(1‐adamantoyl)‐1‐pentylindole, a derivative of JWH‐018 carrying an adamantoyl moiety instead of a naphthoyl group. This finding supports that the listing of synthetic cannabinoids as prohibited substances triggers the appearance of compounds with uncommon substituents. Moreover, it emphasizes the necessity of being aware of the risk of misidentification when using techniques sometimes providing only limited structural information like GC–MS. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

2.
Aryl azides 1 were treated with allenylmagnesium bromide ( 2 ) to generate 1,5‐disubstituted butynyl‐1H‐1,2,3‐triazoles 3 in a domino fashion, which upon CuI‐catalyzed 1,3‐dipolar cycloaddition with aryl azides 4 afforded novel bis‐1H‐1,2,3‐triazoles 5 in quantitative yields (Scheme 1 and Table).  相似文献   

3.
Flavin‐dependent ‘ene’‐reductases (EREDs) are highly selective catalysts for the asymmetric reduction of activated alkenes. This function is, however, limited to enones, enoates, and nitroalkenes using the native hydride transfer mechanism. Here we demonstrate that EREDs can reduce vinyl pyridines when irradiated with visible light in the presence of a photoredox catalyst. Experimental evidence suggests the reaction proceeds via a radical mechanism where the vinyl pyridine is reduced to the corresponding neutral benzylic radical in solution. DFT calculations reveal this radical to be “dynamically stable”, suggesting it is sufficiently long‐lived to diffuse into the enzyme active site for stereoselective hydrogen atom transfer. This reduction mechanism is distinct from the native one, highlighting the opportunity to expand the synthetic capabilities of existing enzyme platforms by exploiting new mechanistic models.  相似文献   

4.
Two new cyclic ‘diarylheptanoids’, myricananone ( 1 ) and myricananadiol ( 2 ), were isolated from the roots of Myrica nana, together with the known compounds myricanol ( 3 ), myricanone ( 4 ), and porson ( 5 ). Their structures were determined by spectroscopic methods, including 1D‐ and 2D‐NMR as well as HR‐ESI‐MS analyses.  相似文献   

5.
The ‘click synthesis’ of some oxiconazole analogs 5a – 5v having 1H‐1,2,3‐triazolyl residues by Huisgen cycloaddition was achieved in four steps (Scheme 1). Oximation of phenacyl chloride ( 1 ) followed by azidation of 2‐chloro‐1‐phenylethanone oxime ( 2 ) provided azido ketoxime 3 . The CuI‐catalyzed Huisgen cycloaddition of 3 with terminal alkynes gave the 4‐substituted (at the triazole) 2‐(1H‐1,2,3‐triazol‐1‐yl)‐1‐phenylethanone oximes 4a – 4i . The O‐alkylation of 4a – 4i with various alkyl halides resulted in the formation of the target molecules 5a – 5v in good yields.  相似文献   

6.
The modification of polymers after the successful achievement of a polymerization process represents an important task in macromolecular science. Cycloaddition reactions, among them the metal catalyzed azide/alkyne ‘click’ reaction (a variation of the Huisgen 1,3‐dipolar cycloaddition reaction between terminal acetylenes and azides) represents an important contribution towards this endeavor. They combine high efficiency (usually above 95%) with a high tolerance of functional groups and solvents, as well as moderate reaction temperatures (25–70 °C). The present review assembles recent literature for applications of this reaction in the field of polymer science (linear polymers, dendrimers, gels) as well as the use of this and related reactions for surface modification on carbon nanotubes, fullerenes, and on solid substrates, and includes the authors own publications in this field. A number of references (>100) are included.

  相似文献   


7.
The reactive 1 : 1 adducts in the reaction between Ph3P and dialkyl acetylenedicarboxylates have been trapped with ‘tosylmethyl isocyanide’ (TsMIC ; 1 ) to yield dialkyl 2‐[(4‐methylphenyl)sulfonyl]‐1H‐pyrrole‐3,4‐dicarboxylates 3 (Scheme 1). The structures of the highly functionalized compounds 3 were corroborated spectroscopically (IR, 1H‐ and 13C‐NMR, and EI‐MS) and by elemental analyses. A plausible mechanism for this type of cyclization is proposed (Scheme 2).  相似文献   

8.
The ‘click synthesis’ of some novel O‐substituted oximes, 7a – 7t , which contain 1,2,3‐triazolediyl residues, as new analogs of β‐adrenoceptor antagonists is described (Schemes 14). The synthesis of these compounds was achieved in four to five steps. The formation of oximes of 9H‐fluoren‐9‐one and benzophenone, i.e., 9a and 9b , respectively, followed by their reaction with propargyl bromide, afforded O‐propargyl oximes 10a and 10b , respectively, which by a subsequent CuI‐catalyzed Huisgen cycloaddition with prepared β‐azido alcohols 11a – 11j (Schemes 2 and 3), led to the target compounds 7a – 7t in good yields.  相似文献   

9.
Electrohydrodynamics is commonly used in microfluidics to control and manipulate the fluid. Though there are studies on the rotation flow in suspended films, the thin film liquid is easily broken and cannot last long hence not applicable in specific applications. Here, we established a three‐dimensional microchamber embedded with two pairs of microelectrodes to investigate the rotational phenomenon of bulk of liquid which we called ‘water fan’ effect based on the electrohydrodynamics force. When proper voltages were applied on these microelectrodes, the tornado‐like rotation would be generated. Both the numerical and experimental results showed that the controllable and continuous rotation could be achieved in the microchamber. In addition, the concentration effect resulting from the rotation flow was also observed. The proposed method offers great promises in providing theoretical and practical guideline in microfluidic devices for mixing, separating, and cooling applications.  相似文献   

10.
The replication of Moloney murine leukaemia virus relies on the formation of a stable homodimeric ‘kissing complex’ of a GACG tetraloop interacting through only two C?G base pairs flanked of 5′‐adjacent unpaired adenosines A9. Previous NMR investigations of a model stem loop 1 has not permitted to reveal the origin of this interaction. Therefore, with the aim of deeper comprehension of the phenomena, the model sequence 10 was prepared where position 9 has been substituted for a nucleoside offering a wider π‐stacking. In this context, the wyosine phosphoramidite building block 2 was prepared and incorporated by adapting the conditions of the automated synthesis and developing original templated enzymatic ligation. However, no ‘kissing interaction’ has been observed for this model sequence 10 due to steric hindrance as confirmed by computational simulation. Consequently, several other model sequences, 18, 23 – 26 , containing modified nucleosides were prepared. Finally, the importance of the cross‐loop H‐bond between G8 and G11 nucleobases was revealed by preparing a 18mer RNA hairpin 27 , where the guanosine G8 has been substituted for inosine. The latter, which does not possess a C3 amino function compared to guanosine, is unable to form any ‘kissing complex’ demonstrating the importance of this secondary interaction in the formation of the complex.  相似文献   

11.
Replacement of the cis‐Me group by CH2F in the imidazolidinone organocatalyst specified in the title (so‐called McMillan generation‐I catalyst) leads to reversal of the product configuration in the title reaction. The topicity reversal in the nucleophilic addition step must arise either from cis‐addition with respect to the benzylic substituent of an (E)‐iminium ion intermediate or from trans‐addition to the corresponding (Z)‐iminium ion. Mechanistic investigations have not provided evidence for either one of these two possibilities, so far.  相似文献   

12.
A new polypyridyl ligand containing a nitro group and two new ruthenium complexes of it were synthesized. The two complexes exhibited non‐dppz DNA ‘light switch’ effects after interaction with calf thymus DNA. Introducing both electron‐withdrawing group (─ NO2) and electron‐donating group (─ CH3) may be the reason that the two complexes display DNA ‘light switch’ behaviors. Furthermore, one of the complexes showed higher photocleavage activity, topoisomerase I inhibition activity and DNA affinity than the other. The present work shows that the more active complex can be employed as a non‐dppz DNA ‘light switch’, DNA photocleaver and topoisomerase I inhibitor. In addition, the two complexes have no or weak cytotoxic activities against Eca‐109 and A549 cells.  相似文献   

13.
Electrospray‐generated precursor ions usually follow the ‘even‐electron rule’ and yield ‘closed shell’ fragment ions. We characterize an exception to the ‘even‐electron rule.’ In negative ion electrospray mass spectrometry (ES‐MS), 2‐(ethoxymethoxy)‐3‐hydroxyphenol (2‐hydroxyl protected pyrogallol) easily formed a deprotonated molecular ion (M‐H)? at m/z 183. Upon low‐energy collision induced decomposition (CID), the m/z 183 precursor yielded a radical ion at m/z 124 as the base peak. The radical anion at m/z 124 was still the major fragment at all tested collision energies between 0 and 50 eV (Elab). Supported by computational studies, the appearance of the radical anion at m/z 124 as the major product ion can be attributed to the combination of a low reverse activation barrier and resonance stabilization of the product ions. Furthermore, our data lead to the proposal of a novel alternative radical formation pathway in the protection group removal of pyrogallol. Copyright © 2009 John Wiley & Sons, Ltd.  相似文献   

14.
The use of chemical warfare agents has become an issue of emerging concern. One of the challenges in analytical monitoring of the extremely toxic ‘V’‐type chemical weapons [O‐alkyl S‐(2‐dialkylamino)ethyl alkylphosphonothiolates] is to distinguish and identify compounds of similar structure. MS analysis of these compounds reveals mostly fragment/product ions representing the amine‐containing residue. Hence, isomers or derivatives with the same amine residue exhibit similar mass spectral patterns in both classical EI/MS and electrospray ionization‐MS, leading to unavoidable ambiguity in the identification of the phosphonate moiety. A set of five ‘V’‐type agents, including O‐ethyl S‐(2‐diisopropylamino)ethyl methylphosphonothiolate (VX), O‐isobutyl S‐(2‐diethylamino)ethyl methylphosphonothiolate (RVX) and O‐ethyl S‐(2‐diethylamino)ethyl methylphosphonothiolate (VM) were studied by liquid chromatography/electrospray ionization/MS, utilizing a QTRAP mass detector. MS/MS enhanced product ion scans and multistage MS3 experiments were carried out. Based on the results, possible fragmentation pathways were proposed, and a method for the differentiation and identification of structural isomers and derivatives of ‘V’‐type chemical warfare agents was obtained. MS/MS enhanced product ion scans at various collision energies provided information‐rich spectra, although many of the product ions obtained were at low abundance. Employing MS3 experiments enhanced the selectivity for those low abundance product ions and provided spectra indicative of the different phosphonate groups. Study of the fragmentation pathways, revealing some less expected structures, was carried out and allowed the formulation of mechanistic rules and the determination of sets of ions typical of specific groups, for example, methylphosphonothiolates versus ethylphosphonothiolates. The new group‐specific ions elucidated in this work are also useful for screening unknown ‘V’‐type agents and related compounds, utilizing precursor ion scan experiments. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

15.
The synthesis of several 18‐membered cyclodepsipeptides with an alternating sequence of α,α‐disubstituted α‐amino acids and α‐hydroxy acids (compounds 14a – 14e ) is described. The ring closure via macrolactonization was accomplished by treatment of a diluted suspension of the corresponding linear precursors 12a – 12e in toluene with HCl gas, i.e., the so‐called ‘direct amide cyclization’. The incorporation of the α,α‐disubstituted α‐amino acids was achieved via the ‘azirine/oxazolone method’ with 2H‐azirin‐3‐amines of type 6 and 9 as building blocks. The structure of the cyclic depsipeptide 14a was established by X‐ray crystallography.  相似文献   

16.
Polycyclic ‘cage’ ketones, such as pentacyclo[5.4.0.02,6.03,10.05,9]undecan‐8‐one ( 10 ), pentacyclo[5.4.0.02,6.03,10.05,9]undecane‐8,11‐dione ( 11 ), and adamantan‐2‐one ( 16 ) were treated with the nucleophilic dimethoxycarbene (DMC; 1 ), which was generated thermally from 2,5‐dihydro‐2,2‐dimethoxy‐5,5‐dimethyl‐1,3,4‐oxadiazole ( 4a ) in boiling toluene. In this ‘one‐pot’ procedure, the α‐hydroxycarboxylic acid ester 12 or a corresponding derivative 15 or 17 was obtained (Schemes 4–7). Additionally, ‘cage’ thione 21 was treated with DMC under the same conditions yielding dimethoxythiirane 22 (Scheme 8). Subsequent hydrolysis or desulfurization (followed by hydrolysis on silica gel) of 22 gave α‐mercaptocarboxylate 25 and the corresponding desulfurized ester 24 , respectively. In all cases, the addition of DMC occurred stereoselectively, and the addition from the exo‐face is postulated to explain the structures of the isolated products.  相似文献   

17.
An atom‐economic Pd0‐catalyzed synthesis of a series of pinacol‐type indolylboronates 3 from the corresponding bromoindole substrates 2 and pinacolborane (pinBH) as borylating agent was elaborated. The optimal catalyst system consisted of a 1 : 2 mixture of [Pd(OAc)2] and the ortho‐substituted biphenylphosphine ligand L‐3 (Scheme 4, Table). Our synthetic protocol was applied to the fast, preparative‐scale synthesis of 1‐substituted indolylboronates 3a – h in the presence of different functional groups, and at a catalyst load of only 1 mol‐% of Pd.  相似文献   

18.
Chemical modification reactions of alkyne containing polyHEMA‐based macroporous network structures (cryogels) by Cu(I) catalyzed azide‐alkyne ‘click’ cycloaddition reactions and their monitoring and quantification with high‐resolution magic angle spinning (hr‐MAS) NMR spectroscopy are reported. Complete conversion is obtained when benzylazide is reacted with the grafted alkyne function, but only partial conversion is observed when using azide‐modified poly(ethylene glycol) (PEG‐N3). Subsequent addition of benzylazide consumes all remaining alkyne groups. All chemical modifications are easily monitored at each stage using hr‐MAS NMR spectroscopy. The alkyne functionality and the resulting triazole ring provide well resolved 1H resonances to monitor and quantify the progress of such ‘click’ reactions in general.

  相似文献   


19.
The reaction of methyl N‐(2,2‐dimethyl‐2H‐azirin‐3‐yl)‐L ‐prolinate ( 2a ) with thiobenzoic acid at room temperature gave the endothiopeptide Bz‐AibΨ[CS]‐Pro‐OMe ( 7 ) in high yield. In an analogous manner, (benzyloxy)carbonyl (Z)‐protected proline was transformed into the thioacid, which was reacted with 2a to give the endothiotripeptide Z‐Pro‐AibΨ[CS]‐Pro‐OMe ( 12 ). The corresponding thioacid of 7 was prepared in situ via saponification, formation of a mixed anhydride, and treatment with H2S. A second reaction with 2a led to the endodithiotetrapeptide 9 , but extensive epimerization at Pro2 was observed. Similarly, saponification of 12 and coupling with either 2a or H‐Phe‐OMe and 2‐(1H‐benzotriazol‐1‐yl)‐1,1,3,3‐tetramethyluronium tetrafluoroborate/1‐hydroxy‐1H‐benzotriazole (TBTU/HOBt) gave the corresponding endothiopeptides as mixtures of two epimers. The synthesis of the pure diastereoisomer BzΨ[CS]‐Aib‐Pro‐AibΨ[CS]‐N(Me)Ph ( 21 ) was achieved via isomerization of 7 to BzΨ[CS]‐Aib‐Pro‐OMe ( 16 ), transformation into the corresponding thioacid, and reaction with N,2,2‐trimethyl‐N‐phenyl‐2H‐azirin‐3‐amine ( 1a ). The structures of 12 and 21 were established by X‐ray crystallography.  相似文献   

20.
The synthesis of four solid solutions (NH4)10–xKx[H2W12O42] · n H2O (x/n: 6.9/9.7, 5.9/9.5, 3.3/8.5, and 2.6/9.0) with the structure of triclinic K10[H2W12O42] · 10 H2O (I) and two solid solutions (x/n: 3.1/4.0, and 2.6/4.0) with the structure of monoclinic (NH4)10[H2W12O42] · 4 H2O (II), was accomplished by a new method from ammoniacal monotungstate solution by adding the appropriate amount of potassium hydroxide and the release of ammonia during evaporative crystallization. The preparation of corresponding single crystals was achieved by slow evaporation of saturated solutions of the corresponding polycrystalline samples according to the method of isothermal evaporation. The study of coordination and space filling behavior of the potassium and ammonium cations, crystal water molecules, and the paratungstate ‘Z' anion revealed that the predominance of the triclinic structure (2.6 ≤ x ≤ 10.0) in the system is effected by the bulkiness of NH4+. The transition area (2.6 ≤ x ≤ 3.1; 4 ≤ n ≤ 8.5) with coexisting triclinic and monoclinic mixed crystals represents the miscibility gap, typical for a heterotypic substitutional solid solution. The ‘resistance' of three specific K+ positions, to be substituted by NH4+, is caused by peculiarities of bond lengths, coordination numbers, and character of coordinated neighbors.  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