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 31 毫秒
1.
Two new ternary tetrazolate Eu(III) complexes with phosphine oxide co‐ligands Eu(PTO)3·(P1/P2) [PTO = 5‐(2‐pyridyl‐1‐oxide)tetrazole, P1 = diphenylphosphorylamino‐phenylphosphoryl‐benzene, P2 = diphenylphosphorylpyridine)‐bis‐isobutyricphosphoryl] were synthesized and characterized using UV, fluorescence, IR and 1H NMR spectroscopic techniques. The analytical data prove that the complexes are mononuclear in nature and the central Eu(III) ion is coordinated by three N and three O atoms of tetrazolate, and two O atoms of the corresponding bidentate phosphine oxide ligands. The ancillary ligand increased the photoluminescence efficiency of Eu(PTO)3·P1 (complex 3) by twofold compared with our previously reported Eu(PTO)3 complex (complex 1). Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

2.
We describe the resolution of a planar chiral cationic iridium complex [Cp*Ir(η5‐2‐methyl‐oxodienyl)][OT f] ( 2 ) following the counterion strategy, where anion metathesis by Δ‐TRISPHAT generates the two diastereomers (pR, pS)‐[Cp*Ir(η5‐2‐methyl‐oxodienyl)][Δ‐TRISPHAT] ( 3a , 3a' ). Upon fractional crystallization both compounds were separated as confirmed by 1H nuclear magnetic resonance (NMR) and circular dichroism studies recorded in solution. The latter represents the key‐complex precursors for the enantioselective synthesis of metallated o‐quinone methide complexes ( 4a , 4a' ). Chirality 25:449–454, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

3.
Lactoferricin (LfB) is a 25‐residue innate immunity peptide released by pepsin from the N‐terminal region of bovine lactoferrin. A smaller amidated peptide, LfB6 (RRWQWR‐NH2) retains antimicrobial activity and is thought to constitute the “antimicrobial active‐site” (Tomita, Acta Paediatr Jpn. 1994; 36 : 585–91). Here we report on N‐acylation of 1‐Me‐Trp5‐LfB6, Cn‐RRWQ[1‐Me‐W]R‐NH2, where Cn is an acyl chain having n = 0, 2, 4, 6 or 12 carbons. Tryptophan 5 (Trp5) was methylated to enhance membrane binding and to allow for selective deuteration at that position. Peptide/lipid interactions of Cn‐RRWQ[1‐Me‐W ]R‐NH2 (deuterated 1‐Me‐Trp5 underlined), were monitored by solid state 31P NMR and 2H NMR. The samples consisted of macroscopically oriented bilayers of mixed neutral (dimyristoylphosphatidylcholine, DMPC) and anionic (dimyristoylphosphatidylglycerol, DMPG) lipids in a 3:1 ratio with Cn‐RRWQ[&1‐Me‐W ]R‐NH2 peptides added at a 1:25 peptide to lipid ratio. 2H‐NMR spectra reveal that the acylated peptides are well aligned in DMPC:DMPG bilayers. The 2H NMR quadrupolar splittings suggest that the 1‐Me‐Trp is located in a motionally restricted environment, indicating partial alignment at the membrane interface. 31P‐NMR spectra reveal that the lipids are predominantly in a bilayer configuration, with little perturbation by the peptides. Methylation alone, in C0‐RRWQ[1‐Me‐W ]R‐NH2, resulted in a 3–4 fold increase in antimicrobial activity against E. coli. N‐acylation with a C12 fatty acid enhanced activity almost 90 fold. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

4.
Boczkowska M  Guga P  Stec WJ 《Biochemistry》2002,41(41):12483-12487
Thermodynamic data regarding the influence of P-chirality on stability of duplexes formed between phosphorothioate DNA oligonucleotides (of either stereo-defined all-R(P) or all-S(P) or random configuration at the P atoms) and complementary DNA or RNA strands are presented. Measured melting temperatures and calculated DeltaG(37)(o) values showed that duplexes formed by PS-DNA oligomers with DNA strands are less stable than their unmodified counterparts. However, relative stability of the duplexes ([all-R(P)]-PS-DNA/DNA vs [all-S(P)]-PS-DNA/DNA) depends on their sequential composition rather than on the absolute configuration of PS-oligos, contrary to the results of theoretical considerations and molecular modeling reported in the literature. On the other hand, for all six analyzed pairs of diastereomers, the [all-R(P)]-PS isomers form more stable duplexes with RNA templates, but the origin of stereodifferentiation depends on the sequence with more favorable entropy and enthalpy factors which correlated with dT-rich and dA/dG-rich PS-oligomers, respectively.  相似文献   

5.
We designed four fluorinated Phe‐incorporated ascidiacyclamide ([Phe]ASC) analogs, (cyclo(?Xxx1‐oxazoline2‐d ‐Val3‐thiazole4‐Ile5‐oxazoline6‐d ‐Val7‐thiazole8‐)), [(4‐F)Phe]ASC (Xxx1: 4‐fluorophenylalanine), [(3,5‐F2)Phe]ASC (Xxx1: 3,5‐difluorophenylalanine), [(3,4,5‐F3)Phe]ASC (Xxx1: 3,4,5‐trifluorophenylalanine) and [(F5)Phe]ASC (Xxx1: pentafluorophenylalanine), to modulate the π‐electron density of the aromatic ring of the Phe residue. X‐ray diffraction analysis, 1H NMR and CD spectra all suggested that the interactions between the benzene ring of the Xxx1 residue and the alkyl groups of oxazoline2 contribute to the stability of the folded structure of these analogs. Substituting fluorines for the hydrogens progressively weakened those interactions through reducing the π‐electron density, thereby mediating transformation from the folded to square structure. As a result, [(F5)Phe]ASC preferred the square form more than the other analogs did. Also contributing to the preference for the square form may be the hindrance of the rotation around the Cα–Cβ bond by the two ortho‐fluoro substituents of [(F5)Phe]ASC. These findings demonstrate that the structure of ASC can be modulated by using fluorine as an electron‐withdrawing group. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
The tris(pyrazolyl)amine ligands: tris[2-(1-pyrazolyl)methyl]amine (tpma), tris [3,5-dimethyl-1-pyrazolyl)methyl]amine (tdma), tris[2-(1-pyrazolyl)ethyl]amine (tpea), tris[2-(3,5-dimethyl-1-pyrazolyl)ethyl]amine (tdea) and bis(pyrazolyl)amine ligands: bis[2-(1-pyrazolyl)ethyl]amine (bpea) and bis[2-(3,5-dimethyl-1-pyrazolyl)ethyl]amine (bdea) react with [RhCl(cod)]2 in presence of NaBF4 (tpma, tdma and bdea) or AgBF4 (tpea, tdea and bpea) to lead to [Rh(cod)L] (BF4) (L=tpma (1), tdma (2), bdea (3), tpea (4), tdea (5) and bpea (6)). These complexes have been characterised by elemental analyses, conductivity, IR, 1H and 13C NMR spectroscopy and liquid mass (with electrospray) spectrometry. The 1H NMR spectra of 1, 2 show the presence of two isomers in solution in a 3:1 ratio (coordination κ2 or κ3 type) in a thermodynamic equilibrium. The steric bulk of cyclo-octa-1,5-diene causes it to prefer the κ2 mode of bonding as majority. Similar to previous published results, complexes 4 and 5 exist in a sole form in solution (probably κ2 isomer). Finally, the complexes 3 and 6 are fluxional. A NMR study shows that this fluxional process is not frozen at 183 K.  相似文献   

7.
Two new 3,5-dimethylpyrazolic derived ligands that are N1-substituted by diamine chains, 1-[2-(diethylamino)ethyl]-3,5-dimethylpyrazole (L1) and 1-[2-(dioctylamino)ethyl]-3,5-dimethylpyrazole (L2) were synthesised. Reaction of the ligands, L1 and L2, with [MCl2(CH3CN)2] yielded [MCl2(L)] (M = Pd(II), Pt(II)) complexes. These complexes were characterised by elemental analyses, conductivity measurements, IR, 1H, 13C{1H} and 195Pt{1H} NMR spectroscopies. The crystal structure of [PdCl2(L1)] was determined by single-crystal X-ray diffraction methods. The structure consists of mononuclear units. The Pd(II) atom is coordinated by a pyrazolic nitrogen, an amine nitrogen and two chlorine atoms in a cis disposition. In this structure, C-H?Cl, C-H?H-C and C-H?C-H intermolecular interactions have been identified.  相似文献   

8.
A novel ligand, 1‐(naphthalen‐2‐yl)‐2‐(phenylsulthio)ethanone was synthesized using a new method and its two europium (Eu) (III) complexes were synthesized. The compounds were characterized by elemental analysis, coordination titration analysis, molar conductivity, infrared, thermo gravimetric analyzer‐differential scanning calorimetry (TGA‐DSC), 1H NMR and UV spectra. The composition was suggested as EuL5 · (ClO4)3 · 2H2O and EuL4 · phen(ClO4)3 · 2H2O (L = C10H7COCH2SOC6H5). The fluorescence spectra showed that the Eu(III) displayed strong characteristic metal‐centered fluorescence in the solid state. The ternary rare earth complex showed stronger fluorescence intensity than the binary rare earth complex in such material. The strongest characteristic fluorescence emission intensity of the ternary system was 1.49 times as strong as that of the binary system. The phosphorescence spectra were also discussed. Copyright © 2014 John Wiley & Sons, Ltd.  相似文献   

9.
Glycogen synthase kinase‐3 (GSK‐3) plays an important regulatory role in various signaling pathways; such as PI3 K/AKT, which is closely related to the occurrence and development of tumors. At present, the most reported active GSK‐3 inhibitors have the same structure: lactam ring or amide structure. To find out the GSK‐3β small molecule inhibitor with novel, safe, efficient and more uncomplicated synthesis method, we analyzed in‐depth reported crystal‐binding patterns of GSK‐3β small molecule inhibitor with GSK‐3β protein, and designed and synthesized 17 non‐reported 3,5‐diamino‐N‐substituted benzamide compounds. Their structures were confirmed by 1H‐NMR, 13C‐NMR, and HR‐MS. The preliminary screening of tumor cytotoxicity of compounds in vitro was detected by MTT, and their structure–activity relationships were illustrated. The results have shown that 3,5‐diamino‐N‐[3‐(trifluoromethyl)phenyl]benzamide ( 4d ) exhibited significant tumor cytotoxicity against human colon cancer cells (HCT‐116) with IC50 of 8.3 μm and showed commendable selectivity to GSK‐3β. In addition, Compound 4d induced apoptosis to some extent and possessed modest PK properties.  相似文献   

10.
N‐[1‐(4‐(4‐fluorophenyl)‐2,6‐dioxocyclohexylidene)ethyl] (Fde) protected amino acids have been prepared and applied in solid‐phase peptide synthesis monitored by gel‐phase 19F NMR spectroscopy. The Fde protective group could be cleaved with 2% hydrazine or 5% hydroxylamine solution in DMF as determined with gel‐phase 19F NMR spectroscopy. The dipeptide Ac‐L ‐Val‐L ‐Val‐NH2 12 was constructed using Fde‐L ‐Val‐OH and no noticeable racemization took place during the amino acid coupling with N,N′‐diisopropylcarbodiimide and 1‐hydroxy‐7‐azabenzotriazole or Fde deblocking. To extend the scope of Fde protection, the hydrophobic nonapeptide LLLLTVLTV from the signal sequence of mucin MUC1 was successfully prepared using Fde‐L ‐Leu‐OH at diagnostic positions. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
The mixtures of room temperature ionic liquid 1‐ethyl‐3‐methylimidazolium trifluoromethanesulfonate ([EMIM]TFO) and water as electrolytes for reduction of CO2 to CO are reported. Linear sweep voltammetry shows overpotentials for CO2 reduction and the competing hydrogen evolution reaction (HER), both of which vary as a function of [EMIM]TFO concentration in the range from 4 × 10?3m (0.006 mol%) to 4869 × 10?3m (50 mol%). A steady lowering of overpotentials up to an optimum for 334 × 10?3m is identified. At 20 mol% and more of [EMIM]TFO, a significant CO2 reduction plateau and inhibition of HER, which is limited by H2O diffusion, is noted. Such a plateau in CO2 reduction correlates to high CO Faraday efficiencies. In case of 50 mol% [EMIM]TFO, a broad plateau spanning over a potential range of 0.58 V evolves. At the same time, the overpotential for HER is increased by 1.20 V when compared to 334 × 10?3m and, in turn, HER is largely inhibited. The Faraday efficiencies for CO and H2 formation feature 95.6% ± 6.8% and 0.5% ± 0.3%, respectively, over a period of 3 h in a separator divided cell. Cathodic as well as anodic electrochemical stability of the electrolyte throughout this time period is corroborated in 1H NMR spectroscopic measurements.  相似文献   

12.
The chemiluminescence intensity of 1,2‐di[3,4,5‐tri(3,4,5‐trihydroxybenzoyloxy)benzoyloxy] benzene increased in the presence of quaternary ammonium ions, such as acetylcholine chloride, choline chloride or benzyltrimethylammonium chloride. The complex of 1,2‐di[3,4,5‐tri(3,4,5‐trihydroxybenzoyloxy)benzoyloxy] benzene with acetylcholine chloride, choline chloride or benzyltrimethylammonium chloride was investigated by 1H‐NMR spectroscopy. The structure of the complex formed from 1,2‐di[3,4,5‐tri(3,4,5‐trihydroxybenzoyloxy)benzoyloxy] benzene with choline chloride was described by an ab initio quantum chemical calculation. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

13.
We designed five ascidiacyclamide analogues [cyclo(‐Xxx1‐oxazoline2‐d ‐Val3‐thiazole4‐l ‐Ile5‐oxazoline6‐d ‐Val7‐thiazole8‐)] incorporating l ‐1‐naphthylalanine (l ‐1Nal), l ‐2‐naphthylalanine (l ‐2Nal), d ‐phenylalanine (d ‐Phe), d ‐1‐naphthylalanine (d ‐1Nal) or d ‐2‐naphthylalanine (d ‐2Nal) into the Xxx1 position of the peptide. The conformations of these analogues were then examined using 1H NMR, CD spectroscopy, and X‐ray diffraction. These analyses suggested that d ‐enantiomer‐incorporated ASCs [(d ‐Phe), (d ‐1Nal), and (d ‐2Nal)ASC] transformed from the folded to the open structure in solution more easily than l ‐enantiomer‐incorporated ASCs [(l ‐Phe), (l ‐1Nal), and (l ‐2Nal)ASC]. Structural comparison of the two analogues containing isomeric naphthyl groups showed that the 1‐naphthyl isomer induced a more stable open structure than the 2‐naphthyl isomer. In particular, [d ‐1Nal]ASC showed the most significant transformation from the folded to the open structure in solution, and exhibited the strongest cytotoxicity toward HL‐60 cells. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
Capillary electrophoresis (CE) allows the observation of the opposite affinities of the enantiomers of (±)‐verapamil [2‐isopropyl‐2,8‐bis(3,4‐dimethoxyphenyl)‐6‐methyl‐6‐azaoctannitrile, VP] toward β‐cyclodextrin (β‐CD) and heptakis(2,3,6‐tri‐O‐methyl)‐β‐CD (TM‐β‐CD). In addition, in the presence of β‐CD in the background electrolyte, longer migration times and lower separation factors were observed compared to TM‐β‐CD. The binding constants of (+)‐ and (−)‐VP with β‐CD and TM‐β‐CD determined using 13C NMR spectroscopy explain the results observed in CE. Electrospray ionization mass spectrometry (ESI‐MS) was used as an alternative technique for the characterization of VP‐CD complexes. Chirality 11:635–644, 1999. © 1999 Wiley‐Liss, Inc.  相似文献   

15.
In this work, three hydrosoluble azocalix[4]arene derivatives, 5-(o-methylphenylazo)-25,26,27-tris(carboxymethoxy)-28-hydroxycalix[4]arene (o-MAC-Calix), 5-(m-methylphenylazo)-25,26,27-tris(carboxymethoxy)-28-hydroxycalix[4]arene (m-MAC-Calix) and 5-(p-methylphenylazo)-25,26,27-tris(carboxymethoxy)-28-hydroxycalix[4]arene (p-MAC-Calix) were synthesized. Their structures were characterized by infrared spectrum (IR), nuclear magnetic resonance spectrum (1H NMR and 13C NMR) and mass spectrum (MS). The interactions between these compounds and bovine serum albumin (BSA) were studied by fluorescence spectroscopy, UV–vis spectrophotometry and circular dichroic spectroscopy. According to experimental results, three azocalix[4]arene derivatives can efficiently bind to BSA molecules and the o-MAC-Calix displays more efficient interactions with BSA molecules than m-MAC-Calix and p-MAC-Calix. Molecular docking showed that the o-MAC-Calix was embedded in the hydrophobic cavity of helical structure of BSA molecular and the tryptophan (Trp) residue of BSA molecular had strong interaction with o-MAC-Calix. The fluorescence quenching of BSA caused by azocalix[4]arene derivatives is attributed to the static quenching process. In addition, the synchronous fluorescence spectroscopy indicates that these azocalix[4]arene derivatives are more accessible to Trp residues of BSA molecules than the tyrosine (Tyr) residues. The circular dichroic spectroscopy further verified the binding of azocalix[4]arene derivatives and BSA.  相似文献   

16.
Two somatostatin analogues, [99mTc]Demotide and [99mTc]Demotate 4, were compared with [99mTc]Demotate 1, a previously reported somatostatin receptor subtype 2 (sst2) targeting tracer. Conjugates were prepared by coupling an open‐chain tetraamine chelator to D ‐Phe1 of [Tyr3]‐octreotide or [Tyr3]‐octreotate, respectively, via a p‐benzylaminodiglycolic acid spacer adopting solid‐phase peptide synthesis techniques. Peptide conjugates were collected in a highly pure form after chromatographic purification. Eventually, [99mTc]Demotide and [99mTc]Demotate 4 were obtained in ~1 Ci/µmol specific activity and >96% purity after labeling under alkaline conditions. Demotide and Demotate 4 exhibited similar high binding affinities for the sst2 expressed in AR4‐2J cells with IC50 values 0.16 and 0.10 nM, respectively. The (radio)metallated analogues [99mTc]Demotide and [99mTc]Demotate 4 showed equally high affinities to the sst2 during saturation binding assays in AR4‐2J cell membranes (Kds 0.08 and 0.07 nM, respectively). During incubation at 37 °C with AR4‐2J cells, the radiopeptides internalized effectively via a receptor‐mediated process, with [99mTc]Demotate 4 exhibiting a faster internalization rate than [99mTc]Demotide. After injection in athymic mice bearing sst2‐expressing AR4‐2J tumors, the radiotracers showed high and specific uptake in the tumor (>25%ID/g at 1 h) and in the sst2–positive organs. However, both [99mTc]Demotide and [99mTc]Demotate 4 showed unfavorably higher background activity, especially in the abdomen, in comparison to [99mTc]Demotate 1 and are, therefore, less suited than [99mTc]Demotate 1 for sst2‐targeted tumor imaging in man. Copyright © 2005 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
Aims: To evaluate the potential of apple pomace (AP) supplemented with rice husk for hyper citric acid production through solid‐state fermentation by Aspergillus niger NRRL‐567. Optimization of two key parameters, such as moisture content and inducer (ethanol and methanol) concentration was carried out by response surface methodology. Methods and Results: In this study, the effect of two crucial process parameters for solid‐state citric acid fermentation by A. niger using AP waste supplemented with rice husk were thoroughly investigated in Erlenmeyer flasks through response surface methodology. Moisture and methanol had significant positive effect on citric acid production by A. niger grown on AP (P < 0·05). Higher values of citric acid on AP by A. niger (342·41 g kg?1 and 248·42 g kg?1 dry substrate) were obtained with 75% (v/w) moisture along with two inducers [3% (v/w) methanol and 3% (v/w) ethanol] with fermentation efficiency of 93·90% and 66·42%, respectively depending upon the total carbon utilized after 144 h of incubation period. With the same optimized parameters, conventional tray fermentation was conducted. The citric acid concentration of 187·96 g kg?1 dry substrate with 3% (v/w) ethanol and 303·34 g kg?1 dry substrate with 3% (v/w) methanol were achieved representing fermentation efficiency of 50·80% and 82·89% in tray fermentation depending upon carbon utilization after 120 h of incubation period. Conclusions: Apple pomace proved to be the promising substrate for the hyper production of citric acid through solid‐state tray fermentation, which is an economical technique and does not require any sophisticated instrumentation. Significance and Impact of the Study: The study established that the utilization of agro‐industrial wastes have positive repercussions on the economy and will help to meet the increasing demands of citric acid and moreover will help to alleviate the environmental problems resulting from the disposal of agro‐industrial wastes.  相似文献   

18.

Aims

This work was performed to characterize new secondary metabolites with neuraminidase (NA) inhibitory activity from marine actinomycete strains.

Methods and Results

An actinomycete strain IFB‐A01, capable of producing new NA inhibitors, was isolated from the gut of shrimp Penasus orientalis and identified as Streptomyces seoulensis according to its 16S rRNA sequence (over 99% homology with that of the standard strain). Repeated chromatography of the methanol extract of the solid‐substrate culture of S. seoulensis IFB‐A01 led to the isolation of streptoseolactone ( 1 ), and limazepines G ( 2 ) and H ( 3 ). The structures of 1 – 3 were determined by a combination of IR, ESI‐MS, 1D (1H and 13C NMR, and DEPT) and 2D NMR experiments (HMQC, HMBC, 1H‐1H COSY and NOESY). Compounds 1 – 3 showed significant inhibition on NA in a dose‐dependent manner with IC50 values of 3·92, 7·50 and 7·37 μmol l?1, respectively.

Conclusions

This is the first report of two new ( 1 and 2 ) and known ( 3 , recovered as a natural product for the first time in the work) NA inhibitors from the marine‐derived actinomycete S. seoulensis IFB‐A01.

Significance and Impact of the Study

The three natural NA inhibitors maybe of value for the development of drug(s) necessitated for the treatment of viral infections.  相似文献   

19.
A series of new chiral molecular tweezers, di‐(R,R)‐1‐[10‐(1‐hydroxy‐2,2,2‐trifluoroethyl)‐9‐anthryl]‐2,2,2‐trifluoroethyl phthalate (2), isophthalate (3) and terephthalate (4), were synthesized and their structure studied by NMR and molecular mechanics. Their effectiveness as chiral solvating agents for the determination of the enantiomeric purity of chiral compounds using NMR was demonstrated. Chirality 2010. © 2009 Wiley‐Liss, Inc.  相似文献   

20.
A procedure for the synthesis of a11C‐labeled oligopeptide containing [1‐11C]1,2,3,4‐tetrahydro‐β‐carboline‐3‐carboxylic acid ([1‐11C]Tpi) from the corresponding Trp?HCl‐containing peptides has been developed involving a Pictet‐Spengler reaction with [11C]formaldehyde. The synthesis of [1‐11C]Tpi from Trp and [11C]formaldehyde was examined as a model reaction with the aim of developing a facile and effective method for the labeling of peptides with carbon‐11. The Pictet‐Spengler reaction of Trp and [11C]formaldehyde in acidic media (TsOH or HCl) afforded the desired [1‐11C]Tpi in a moderate radiochemical yield. Herein, the application of a Pictet‐Spengler reaction to an aqueous solution of Trp?HCl gave the desired product with a radiochemical yield of 45.2%. The RGD peptide cyclo[Arg‐Gly‐Asp‐D‐Tyr‐Lys] was then selected as a substrate for the labeling reaction with [11C]formaldehyde. The radiolabeling of a Trp?HCl‐containing RGD peptide using the Pictet‐Spengler reaction was successful. Furthermore, the remote‐controlled synthesis of a [1‐11C]Tpi‐containing RGD peptide was attempted by using an automatic production system to generate [11C]CH3I. The radiochemical yield of the [1‐11C]Tpi‐containing RGD at the end of synthesis (EOS) was 5.9 ± 1.9% (n = 4), for a total synthesis time of about 35 min. The specific activity was 85.7 ± 9.4 GBq/µmol at the EOS. Copyright © 2013 European Peptide Society and John Wiley & Sons, Ltd.  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