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 46 毫秒
1.
Hirudin-1 is a highly selective inhibitor of thrombin secreted by the salivary glands of the medicinal leech Hirudo medicinalis. This direct anticoagulant is used for the treatment and prevention of disorders in the blood coagulation system. Apart from the existing recombinant analogue of hirudin-1 (63-desulfato-hirudin-1) its modified analogues possessing higher activity and stability are of medical value. In this study artificial genes of hirudin-1 and two its analogues ([Leu1, Thr2]-hirudin-1 and [Leu1, Thr2]-hirudin-1/3) were synthesized and cloned in an expression vector pTWIN1 in frame with the gene of mini-intein DnaB from Synechocystis sp. Producing strains of the corresponding fusion proteins were constructed using E. coli strain ER2566. Biotechnological schemes for the production of 63-desulfatohirudin-1 and its analogues were developed. The scheme includes the following stages: isolation of the fusion protein, renaturation of the target protein incorporated into the fusion protein, pH-inducible cleavage of the fusion protein, and chromatographic purification of the target product. Antithrombotic activity of the peptides obtained was determined by a standard amidolytic assay. The developed methods for the production of 63-desulfatohirudin-1, [Leu1, Thr2]-63-desulfatohirudin-1 and [Leu1, Thr2]-63-desulfatohirudin-1/3 allowed us to obtain these peptides with high yields (14, 25 and 24 mg per 1 L of cell culture, respectively) and high activity (13423, 33333 and 19802 ATU/mg, respectively).  相似文献   

2.
Occurrence of cyclosporins and cyclosporin-like peptolides in fungi   总被引:1,自引:0,他引:1  
Apart from 17 previously listed fungal taxa producing cyclosporin A and its natural congeners B to Z, several additional and taxonomically diverse strains producing the single novel component [Thr2, Leu5, Leu10]cyclosporin or cyclosporin-like peptolides (eg SDZ 214-103=[Thr2, Leu5, D-Hiv8, Leu10]cyclosporin) have recently been described. We report here the isolation of two further and novel cyclosporins, [Thr2, Leu5, Ala10]cyclosporin (2) and [Thr2, Ile5] cyclosporin (3), from strains classified asAcremonium luzulae (Fuckel) W Gams andLeptostroma anamorph ofHypoderma eucalyptii Cooke & Harkn, respectively. In both new strains the usual pattern of cyclosporins A to Z is not found. The structure elucidations of2 and3 are based on NMR spectroscopy, and biological data (immunosuppressive activity, cyclophilin-binding affinity and antifungal effects) are presented.  相似文献   

3.
The synthesis is described of the protected hexadecapeptide corresponding to the C-terminal sequence 93–108 of baker's yeast iso-I-cytochrome c. The cysteine residue in position 107 of the natural sequence has been substituted by a threonine residue. The rationale of this substitution as well as the synthetic route to the preparation of the hexadecapeptide derivative is discussed.  相似文献   

4.
Proteolysis of semax (Met-Glu-His-Phe-Pro-Gly-Pro, Sem) and its analogues with the substitution of Ala, Gly, Thr, or Trp for the N-terminal Met was studied. This substitution was shown to change the degradation rate of these peptides by leucine aminopeptidase (EC 3.4.11.2, Sigma, Type VI, 9.2 activity units/mg). [Ala1]Sem, [Gly1]Sem, and [Thr1]Sem (the semax analogues) proved to be more stable to the proteolysis than semax itself. It was demonstrated that the primary product of the proteolysis was His-Phe-Pro-Gly-Pro (Sem-5). In the case of [Trp1]Sem, the comparable amount of Glu-His-Phe-Pro-Gly-Pro (Sem-6) was found to be formed along with Sem-5. In was established that all the studied semax analogues could be used as inhibitors of its proteolysis.  相似文献   

5.
The trypsin-catalyzed coupling of bovine (Boc)2-desoctapeptide (B23-B30)-insulin with synthetic octapeptides, H-Gly-X2-X3-X4-Thr-Pro-Lys(Boc)-Thr-OH (X2 = Phe or Ala, X3 = Phe or Ala, X4 = Tyr or Ala), followed by deprotection and purification produced the [AlaB24, ThrB30]-, [AlaB25, ThrB30]-, and [AlaB26, ThrB30]-analogs of bovine insulin in yields of 32, 35, and 32%, respectively. The biological activity of these analogs decreased in the order, normal insulin ([ThrB30]-bovine insulin) = AlaB26-insulin > AlaB25-insulin > AlaB24-insulin, as assayed for receptor binding and some other biological effects, in contrast with the corresponding Leu-analogs of human insulin, in which the activity decreased in the order, normal insulin > LeuB24-insulin > LeuB25-insulin. The affinity to insulin antibodies greatly diminished in both AlaB24-insulin and LeuB24-insulin but not in the B25-substituted analogs. The CD spectra of the Leu- and the Ala-analogs were compared with those of normal insulins to show that no apparent correlation seems to exist between the decrease in biological activity and the conformational changes observed in solution. The effects of organic solvents on the peptide-bond equilibrium and on the stability of trypsin are also discussed.  相似文献   

6.
We describe the solution (1H-nmr) and calculated conformations of the opiatelike peptide dermorphin and the analysis of structure–conformation–activity relationships in the series [Alan]-dermorphin. We used 1H-nmr spectroscopy to study dermorphin and its analogs [Alan]-dermorphin (with n = 1, 2…7) dissolved in dimethylsufoxide. Conformational energy calculations using semiempirical partitioned energy function methods were then carried out on dermorphin and its [L -Ala2]-analog. Agreement between calculation and experiment is satisfying, both suggesting predominance of a type I β-turn around Pro6-Ser7 at the C-terminus and of an extended structure in the central sequence Phe3-Gly4-Tyr5. Detailed analysis by step-by-step substitutions with Ala indicates that intraresidue interactions dominate over medium-range interactions (between adjacent residues), although the latter may also have a noticeable influence in shaping conformations. As a general feature, the effects of substitutions on the arrangement of side chains are always larger on the succeeding residue than on the preceding residue. Almost all the variations of activity observed in the analogs can be explained from conformational changes occurring in the aromatic side chains of the biologically important Tyr1, Phe3, and Tyr5 on substitutions effected on adjacent residues (fluctuations via medium-range interactions).  相似文献   

7.
Syntheses are described of the nociceptin (1–13) amide [NC(1–13)-NH2] and of several analogues in which either one or both the phenylalanine residues (positions 1 and 4), the arginine residues (positions 8 and 12) and the alanine residues (positions 7 and 11) have been replaced by N-benzyl-glycine, N-(3-guanidino-propyl)-glycine and β-alanine, respectively. The preparation is also described of NC(1–13)-NH2 analogues in which either galactose or N-acetyl-galactosamine are β-O-glycosidically linked to Thr5 and/or to Ser10. Preliminary pharmacological experiments on mouse vas deferens preparations showed that Phe4, Thr5, Ala7 and Arg8 are crucial residues for OP4 receptor activation. Manipulation of Phe1 yielded peptides endowed with antagonist activity but [Nphe1] NC(1–13)-NH2 acted as an antagonist still possessing weak agonist activity. Introduction of the βAla residue either in position 7 or 11 of the [Nphe1] NC(1–13)-NH2 sequence, abolished any residual agonist activity and [Nphe1, βAla7] NC(1–13)-NH2 and [Nphe1, βAla11] NC(1–13)-NH2 acted as competitive antagonists only. Modification of both Ala7 and Ala11 abolished the antagonist activity of [Nphe1]NC(1–13)-NH2 probably by hindering receptor binding. Changes at positions 10 and 11 gave analogues still possessing agonist activity. [Ser(βGal)10] NC(1–13)-NH2 displayed an activity comparable with that of NC(1–13)-NH2, [Ser(βGalNAc)10] NC(1–13)-NH2 and [βAla11] NC(1–13)-NH2 were five and 10 times less active, respectively.The α-amino acid residues are of the l-configuration. Standard abbreviations for amino acid derivatives and peptides are according to the suggestions of the IUPAC-IUB Commission on Biochemical Nomeclature (1984), Eur. J. Biochem. 138, 9–37. Abbreviations listed in the guide published in (2003), J. Peptide Sci. 9, 1–8 are used without explanation.  相似文献   

8.
The ability of glucagon and several of its semi-synthetic analogues to stimulate glucose production in isolated rat hepatocytes was measured and compared for relative potencies. The order of decreasing biological activities of glucagon in this assay was as follows: glucagon > [HArg12]-glucagon > [des-Asn28, Thr29][homoserinehydrazide27]-glucagon approx. equal to [des-His1]-glucagon > [des-Asn28, Thr29] [homoserinelactone27]-glucagon > [des-Asn28, Thr29]-[n-butylhomoserineamide27]-glucagon > glucagon1?21. Qualititatively, these results are similar to those obtained previously in the hepatic plasma membrane adenylate cylase assay. Minor exceptions were noted for the hydrazide derivative nd the partial agonist [des-His1]-glucagon, both of which were slightl y more potent relative to glucagon in the glycogenolytic assay than in the adenylate cyclase assay. The assay provides important insight into glucagon structure-function relationships.  相似文献   

9.
Neb-colloostatin (SIVPLGLPVPIGPIVVGPR), an insect oostatic factor found in the ovaries of the flesh fly Neobellieria bullata, strongly induces apoptosis in insect haemocytes. To explain the role of Ser1 and Pro4 residues of Neb-colloostatin in the pro-apoptotic activity of this peptide, the synthesis of a series of analogs was performed, such as: [Ac-Ser1]- (1), [d-Ser1]- (2), [Thr1]- (3), [Asp1]- (4), [Glu1]- (5), [Gln1]- (6), [Ala1]- (7), [Val1]- (8), [d-Pro4]-(9), [Hyp4]- (10), [Acp4]- (11), [Ach4]- (12), [Ala4]- (13), [Ile4]- (14), and [Val4]-colloostatin (15). All peptides were bioassayed in vivo for the pro-apoptotic action on haemocytes of Tenebrio molitor. Additionally, the structural properties of Neb-colloostatin and its analogs were examined by the circular dichroism in water and methanol. Peptides 1, 4, 5, 7, 8, 10, 12, 14, and 15 strongly induce T. molitor haemocytes to undergo apoptosis and they show about 120–230% of the Neb-colloostatin activity at a dose of 1 nM. The CD conformational studies show that the investigated peptides seem to prefer the unordered conformation.  相似文献   

10.
We report the conformational analysis by 1H‐NMR in DMSO and computer simulations involving distance geometry and molecular dynamics simulations of peptoid analogs of the cyclic hexapeptide c‐[Phe11‐Pro6‐Phe7‐d ‐Trp8‐Lys9‐Thr10] L‐363,301 (the numbering refers to the positions in native somatostatin). The compounds c‐[Phe11‐Nphe6‐Nal7‐d ‐Trp8‐Lys9‐Thr10] ( Nphe 6 ‐ Nal 7 analog 1 ), c‐[Nal11‐Nphe6‐Phe7‐d ‐Trp8‐Lys9‐Thr10] ( Nal 11 ‐ Nphe 6 analog 2 ) and c‐[Phe11‐Nnal6‐Phe7‐d ‐Trp8‐Lys9‐Thr10] ( Nnal 6 analog 3 ), where Nphe denotes N‐benzylglycine and Nnal denotes N‐(1‐naphthylmethyl)glycine, are subjected to SAR studies in order to investigate the influence of the bulky naphthyl aromatic ring on the conformation. The Nal 11 ‐ Nphe 6 and Nphe 6 ‐Nal 7 analogs exhibit potent binding to the hsst2, hsst3 and hsst5 receptors, whereas the Nnal 6 analog has decreased binding affinity to all receptors but is more selective towards the hsst2 than the other two analogs and L‐363,301. The conformational search employing distance geometry, energy minimization and molecular dynamic simulations gives insight into the conformational flexibility of these analogs. The molecules adopt both cis and trans orientations of the peptide bond between residues 11 and 6. The cis isomers of these analogs adopt type II′ β‐turns with d ‐Trp in the i+1 position and type VIa β‐turns with the cis peptide bond between residues 6 and 11. The results of free and distance restrained molecular dynamics simulations at 300 K indicate that the Nphe 6 ‐Nal 7 and Nal 11 ‐Nphe 6 compounds adopt a preferred backbone conformation which can be described as ‘folded’ about residues 7 and 10. The Nnal 6 analog, which binds less effectively to the hsst receptors, has a more flexible backbone structure than the Nal 11 ‐Nphe 6 and Nphe 6 ‐Nal 7 analogs and prefers a ‘flat’ structure with regard to the orientations about Phe7 and Thr10 during molecular dynamics simulations. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
Abstract: δ-Opioids mobilize Ca2+ from intracellular stores in undifferentiated NG108-15 cells, but the mechanism involved remains unclear. Therefore, we examined the effect of [d -Pen2,5]enkephalin on inositol 1,4,5-trisphosphate formation in these cells. [d -Pen2,5]enkephalin caused a dose-dependent (EC50 = 3.1 nM) increase in inositol 1,4,5-trisphosphate formation (measured using a specific radioreceptor mass assay), which peaked (25.7 ± 1.2 pmol/mg of protein with 1 µM, n = 9) at 30 s and returned to basal levels (10.6 ± 0.9 pmol/mg of protein, n = 9) within 4–5 min. This response was fully naloxone (1 µM) reversible and pertussis toxin (100 ng/ml for 24 h) sensitive. Preincubation with Ni2+ (2.5 mM) or nifedipine (1 µM) had no effect on the [d -Pen2,5]enkephalin (1 µM)-induced inositol 1,4,5-trisphosphate response, and K+ (80 mM) was unable to stimulate inositol 1,4,5-trisphosphate formation, indicating Ca2+ influx-induced activation of phospholipase C is not involved. Preincubation with the protein kinase C inhibitor Ro 31-8220 (1 µM) enhanced, whereas acute exposure to phorbol 12,13-dibutyrate (1 µM) abolished, the [d -Pen2,5]enkephalin (0.1 µM)-induced inositol 1,4,5-trisphosphate response, suggesting protein kinase C exerts an autoinhibitory feedback action. [d -Pen2,5]Enkephalin also dose-dependently (EC50 = 2.8 nM) increased the intracellular [Ca2+], which was maximal (24 nM increase with 1 µM, n = 5) at 30 s. This close temporal and dose-response relationship strongly suggests that δ-opioid receptor-mediated increases in intracellular [Ca2+] results from inositol 1,4,5-trisphosphate-induced Ca2+ release from intracellular stores, in undifferentiated NG108-15 cells.  相似文献   

12.
Abstract Trimeresurus flavoviridis snakes inhabit the southwestern islands of Japan. A phospholipase A2 (PLA2), named PL-Y, was isolated from Okinawa T. flavoviridis venom and its amino acid sequence was determined from both protein and cDNA. PL-Y was unable to induce edema. In contrast, PLA-B, a PLA2 from Tokunoshima T. flavoviridis venom, which is different at only three positions from PL-Y, is known to induce edema. A new PLA2, named PLA-B′, which is similar to PLA-B, was cloned from Amami-Oshima T. flavoviridis venom gland. Three T. flavoviridis venom basic [Asp49]PLA2 isozymes, PL-Y (Okinawa), PLA-B (Tokunoshima), and PLA-B′ (Amami-Oshima), are identical in the N-terminal half but have one to four amino acid substitutions in the β1-sheet and its vicinity. Such interisland sequence diversities among them are due to isolation in the different environments over 1 to 2 million years and appear to have been brought about by natural selection for point mutation in their genes. Otherwise, a major PLA2, named PLA2, ubiquitously exists in the venoms of T. flavoviridis snakes from the three islands with one to three synonymous substitutions in their cDNAs. It is assumed that the PLA2 gene is a prototype among T. flavoviridis venom PLA2 isozyme genes and has hardly undergone nonsynonymous mutation as a principal toxic component. Phylogenetic analysis based on the amino acid sequences revealed that T. flavoviridis PLA2 isozymes are clearly separated into three groups, PLA2 type, basic [Asp49]PLA2 type, and [Lys49]PLA2 type. Basic [Asp49]PLA2-type isozymes may manifest their own particular toxic functions different from those of the isozymes of the PLA2 type and [Lys49]PLA2 type.  相似文献   

13.
We report details of the chemical synthesis of two fragments reproducing the C-terminal sequences 71-108 and 70-108 of Saccharomices cerevisiae cytochrome c. Preparation of the fragments employed classical solution methods and a fragment-condensation strategy; they have been used, together with a third fragment (sequence 67-108) [L. Moroder, B. Filippi, G. Borin & F. Marchiori (1975) Biopolymers 14 , 2061–2074], in the semisynthesis of chimeric cytochromes [C. J. A. Wallace, G. Corradin, F. Marchiori & G. Borin (1986) Biopolymers 25 , 2121–2132].  相似文献   

14.
The C-terminal region of parathyroid hormone-related protein (PTHrP) containing the sequence (107–111) appears to be a potent inhibitor of osteoclastic bone resorption. In the present study, we have investigated the effect of human (h)PTHrP (107–139) and hPTHrP (107–111)NH2 on the proliferation of osteoblastic rat osteosarcoma UMR 106 cells. We found that both C-terminal PTHrP peptides, like hPTHrP (1–141), were antimitogenic for these cells, between 1 pM and 10 nM. [Tyr34]hPTHrP (1–34)NH2 was as potent as these peptides but less effective as growth inhibitor in these cells. UMR 106 cells were found to produce and secrete immunoreactive PTHrP. Addition of anti-PTHrP neutralizing antibodies to C- and N-terminal epitopes of PTHrP increased the growth of these cells. Our data suggest that the antiproliferative effect of these C-terminal PTHrP analogs may be independent of cyclic adenosine 3′:5′-monophosphate (cAMP) and mediated by protein kinase C. These findings support an autocrine role of PTHrP in bone metabolism. J. Cell. Physiol. 170:209–215, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

15.
Earlier studies have demonstrated that a high (mM) extracellular Ca2+ concentration triggers intracellular [Ca2+] signals with a consequent inhibition of bone resorptive activity. We now report that micromolar concentrations of the divalent cation, Ni2+, elicited rapid and concentration-dependent elevations of cytosolic [Ca2+]. The peak change in cytosolic [Ca2+] increased monotonically with the application of [Ni2+] in the 50–5,000 μM range in solutions containing 1.25 mM-[Ca2+] and 0.8 mM-[Mg2+]. The resulting concentration-response function suggested Ni2+-induced activation of a single class of binding site (Hill coefficient = 1). The triggering process also exhibited a concentration-dependent inactivation in which conditioning Ni2+ applications in the range 5–1,500 μM-[Ni2+] inhibited subsequent responses to a maximally effective [Ni2+] of 5,000 μM. Ni2+-induced cytosolic [Ca2+] responses were not dependent on extracellular [Ca2+]. Thus, when 5,000 μM-[Ni2+] was applied to osteoclasts in Ca2+-free, ethylene glycol bis-(aminoethyl ether) tetraacetic acid (EGTA)-containing medium (≤5 nM-[Ca2+] and 0.8 mM-[Mg2+]), cytosolic [Ca2+] responses resembled those obtained in the presence of 1.25 mM-[Ca2+]. Prior depletion of intracellular Ca2+ stores by ionomycin prevented Ni2+-induced cytosolic [Ca2+] responses, suggesting a major role for intracellular Ca2+ redistribution in the response to Ni2+. The effects of Ni2+ were also modulated by the extracellular concentration of the divalent cations, Ca2+ and Mg2+. When these cations were not added to the culture medium (0 μM-[Ca2+] and [Mg2+]), even low [Ni2+] ranging between 5 pM and 50 μM elicited progressively larger cytosolic [Ca2+] transients. However, the response magnitude decreased at higher, 250–5,000 μM-[Ni2+], resulting in a “hooked” concentration-response curve. Furthermore, increasing extracellular [Mg2+] or [Ca2+] (0–1 mM) diminished the response to 50 μM-[Ni2+], a concentration on the rising phase of the “hook.” Similar increases (0–10 mM) in extracellular [Mg2+] or [Ca2+] increased the response to 5,000 μM-[Ni2+], a concentration on the falling phase of the “hook”. These findings are consistent with the existence of a membrane receptor strongly sensitive to Ni2+ as well as the divalent cations, Ca2+ and Mg2+. Receptor occupancy apparently activates intracellular Ca2+ release followed by inactivation. Furthermore, repriming is independent of intracellular Ca2+ stores, suggesting that such inactivation operates at a transduction step between receptor occupancy and intracellular Ca2+ release. © 1993 Wiley-Liss, Inc.  相似文献   

16.
1H-nmr studies of [pGlu6]SP6–11, [gpGlu6,mPhe7]SP6–11, and [pGlu6,N-CH3Phe7]SP6–11 in DMSO-d6 reveal characteristic chemical shifts, 3JNH-αCH, temperature dependence, as well as deuterium exchange half-times. Marked similarities are revealed for the two first analogs, whereas the N-methylated analog is clearly different. Possible conformations are considered.  相似文献   

17.
Abstract: The regulation of adenylate cyclase activity by adrenocorticotropin/α-melanocyte–stimulating hormone (ACTH/MSH)-like peptides was investigated in rat brain slices using a superfusion method. Adenylate cyclase activity was concentration-dependently increased by ACTH-(1–24), α-MSH (EC50 values 16 and 6 nM, respectively), and [Nle4,D-Phe7]α-MSH (EC50 value 1.6 nM), in the presence of forskolin (1 μM, optimal concentration). 1-9-Dideoxy-forskolin did not augment the response of adenylate cyclase to ACTH-(1–24). Various peptide fragments were tested for their ability to enhance [3H]cyclic AMP production. [Nle4,D-Phe7]α-MSH increased [3H]cyclic AMP formation with a maximal effect of 30% and was more potent than ACTH-(1–24), ACTH-(1–16)-NH2, α-MSH, ACTH-(1–13)-NH2, [MetO4]α-MSH, [MetO24,D-Lys8,Phe9]ACTH-(4–9), ACTH-(7–16)-NH2, ACTH-(1–10), and ACTH-(11–24), in order of potency. This structure–activity relationship resembles that found for the previously described peptide-induced display of excessive grooming. ACTH-(1–24) stimulated adenylate cyclase activity in both striatal (maximal effect, ?20%) and septal slices (maximal effect, ?40%), but not in hippocampal or cortical slices. Lesioning of the dopaminergic projections to the striatum did not result in a diminished effect of [Nle4,D-Phe7]α-MSH on [3H]cyclic AMP accumulation, which indicates that the ACTH/MSH receptor–stimulated adenylate cyclase is not located on striatal dopaminergic terminals. ACTH-(1–24) did not affect the dopamine D1 or D2 receptor–mediated modulation of adenylate cyclase activity. Based on the present data, we suggest that the binding of endogenous ACTH or α-MSH to a putative ACTH/MSH receptor in certain brain regions leads to the activation of a signal transduction pathway using cyclic AMP as a second messenger.  相似文献   

18.
All seven possible bradykinin (BK) analogs containing Aib in place of proline have been synthesized by the solid phase method and assayed for in vitro myotropic activity on the guinea pig ileum and rat uterus, and in vivo on the rat blood pressure, both by intravenous and intra-aortic administration. [Aib2,3]-BK, [Aib2,7]-BK, and [Aib2,3,7]-BK had no in vivo or in vitro activities; [Aib2]-BK, [Aib3]-BK and [Aib3,7]-BK had moderate BK-like activities and a significantly increased resistance to pulmonary inactivation in the rat ([Aib3,7]-BK was totally resistant). [Aib7]-BK was found to be the most active position seven BK analog yet assayed on the rat blood pressure, and shows remarkably high ileum (4 times BK) and intravenous rat blood pressure (6 times BK) activity.  相似文献   

19.
Polymers based on thieno[3,4‐c]pyrrole‐4,6‐dione derivatives are interesting and promising candidates for organic bulk heterojunction solar cells. Herein, a series of push–pull conjugated polymers based on thieno[3,4‐c]pyrrole‐4,6‐dione (TPD), furo[3,4‐c]pyrrole‐4,6‐dione (FPD), and selenopheno[3,4‐c]‐pyrrole‐4,6‐dione (SePD) have been synthesized by direct heteroarylation polymerization and fully characterized. The impacts of both the heteroatom (sulfur, oxygen, and selenium) and the side chain (branched or linear) of [3,4‐c]pyrrole‐4,6‐dione unit on the electro‐optical properties have been investigated. Among polymers developed, two new highly processable terthiophene–SePD ( P4 ) and dithienosilole–SePD ( P9 ) copolymers led to air‐processed polymer solar cells with power conversion efficiencies of 5.1% and 7.1% using the following inverted configuration: ITO/ZnO/Polymer:PCBM/MoO3/Ag. These promising results make P4 and P9 good candidates for further upscaling and device optimization.  相似文献   

20.
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