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 390 毫秒
1.
It is important for ion channel peptides to have energetic stability and ion-selectivity for development of some medicines. In the present study, our objective was to achieve formation of energetically stable and ion-selective channels in the membrane using cyclic tetrapeptides. We succeeded in formation of energetically stable and ion-selective channels using two cyclic tetrapeptides cyclo(D-Ala-Dap)(2) (Dap; l-2,3-diaminopropionic acid) and cyclo(D-Ala-Glu)(2). The results of ion channel recording suggested that the cationic cyclo(D-Ala-Dap)(2) was resulted in Cl(-) anion-selective and the anionic cyclo(D-Ala-Glu)(2) led to K(+) cation-selective ion channel formation, respectively. This ion selectivity may be attributed to the charge state of peptides. And a low-hydrophobic cyclic tetrapeptide; cyclo(D-Ala-Dap)(2) had a tendency to form stable ion channel compared to more high-hydrophobic ones; cyclo(D-Phe-Lys)(2), cyclo(D-Phe-Dap)(2) and cyclo(D-Ala-Lys)(2). Our findings will shed light on the field of ion channel peptide study, especially cyclic one.  相似文献   

2.
从中国民间草药短瓣花(Brachystemma calycinum D.Don)干燥根的乙醇提取物中分离得到4个新的微量环八肽,命名为短瓣花环肽A、B、C和D。经光谱分析鉴定。它们的结构分别为cyclo(Pro^1-Phe-Leu-Ala^1-Thr-Pro^2-Ala^2-Gly)(1)、cyclo(Pro^1-Ala-Phe-Trp-Asp-Pro^2-Leu-Gly)(2)、cyclo(Pro^1-Ⅱe-Gly-Pro^2-Val-Ala^1-Ala^2-Tyr)(3)和cyclo(Pro-OMet-Trp-Ⅱe-Gly-Ala-Leu-Asp)(4)。  相似文献   

3.
The three-dimensional structure of a potent SSTR3-selective analogue of somatostatin, cyclo(3-14)H-Cys(3)-Phe(6)-Tyr(7)-D-Agl(8)(N(beta) Me, 2-naphthoyl)-Lys(9)-Thr(10)-Phe(11)-Cys(14)-OH (des-AA(1, 2, 4, 5, 12, 13)[Tyr(7), D-Agl(8)(N(beta) Me, 2-naphthoyl)]-SRIF) (peptide 1) has been determined by (1)H NMR in water and molecular dynamics (MD) simulations. The peptide exists in two conformational isomers differing mainly by the cis/trans isomerization of the side chain in residue 8. The structure of 1 is compared with the consensus structural motifs of other somatostatin analogues that bind predominantly to SSTR1, SSTR2/SSTR5 and SSTR4 receptors, and to the 3D structure of a non-selective SRIF analogue, cyclo(3-14)H-Cys(3)-Phe(6)-Tyr(7)-D-2Nal(8)-Lys(9)-Thr(10)-Phe(11)-Cys(14)-OH (des-AA(1, 2, 4, 5, 12, 13)[Tyr(7), D-2Nal(8)]-SRIF) (peptide 2). The structural determinant factors that could explain selectivity of peptide 1 for SSTR3 receptors are discussed.  相似文献   

4.
In the present study, novel eight GS derivatives having the octanoyl-(Lys)(n)- moieties, cyclo{-Val-Orn-Leu-d-Phe-Pro(4β-NH-X)-Val-Orn-Leu-d-Phe-Pro-} {X=-H (1), and -(Lys)(n)-CO(CH(2))(6)CH(3)n=0 (2), 1 (3), 2 (4), and 3 (5)} and cyclo{-Val-Orn-Leu-d-Phe-Pro(4α-NH-X)-Val-Orn-Leu-d-Phe-Pro-} {X=-H (6), and -(Lys)(n)-CO(CH(2))(6)CH(3)n=1 (7), and 2 (8)} were synthesized. Among them, 4, 5 and 8 result the high antibiotic activity against both Gram-positive and Gram-negative microorganisms tested. In addition, 4 and 5 showed very low hemolytic activity compared with that of GS. Thus, the introduction of the excess amino groups and the fatty acyl moiety to the γ-NH(2) group of Pro(5) residue in GS molecule lowered the unwanted hemolytic activity and enhanced the desired antibiotic activity.  相似文献   

5.
Two new cyclic peptides, named sajaponicin C (1) and sajaponicin D (2), were isolated from the whole plants of Sagina japonica (Caryophyllaceae). Their structures were determined as cyclo(Pro(2)-Leu(2)-Tyr-Leu(1)-Phe(1)-Pro(3)-Phe(2)-Pro(1)) (1) and cyclo(Pro(1)-Pro(2)-Pro(3)-Pro(4)-Phe(1)-Gly-Thr-Ser-Phe(2)-Ile-Tyr) (2) on the basis of spectroscopic data, especially by two-dimensional (2D) NMR techniques.  相似文献   

6.
Re-investigation of the culture broth of the marine bacterium Staphylococcus sp. (no. P-100826-4-6) afforded a new cyclic tetrapeptide namely staphylopeptide A (1), along with five known compounds, cyclo (l-prolyl-l-valine) (2), cyclo (l-prolyl-l-tyrosine), (3), cyclo (l-prolyl-l-alanine) (4), l-phenylalanine (5), and l-tryptophan (6). The structures of the isolated compounds were determined by extensive IR, 1D (1H and 13C) and 2D (1H-1H COSY, HSQC, HMBC, and NOESY) NMR and HRFABMS spectral measurements. The antimicrobial activity of the isolated compounds (1–6) was evaluated.  相似文献   

7.
Biologically active cyclic tetrapeptides, usually found among fungi metabolites, exhibit phytotoxic or cytostatic activities that are likely to be governed by specific conformations adopted in solution. For conformational studies and drug design, there is a strong interest in using fast and reliable methods to determine correctly the conformational population of cyclotetrapeptides. We show here that standard molecular mechanics computational approach gives satisfactory results. The method was validated step by step by experimental data either obtained after synthesis and NMR analysis, or found in the literature. The cyclo(Gly)(4), cyclo(Ala)(4), cyclo(Sar)(4), and cyclo(SarGly)(2) peptides were used to evaluate the prediction of the peptide backbone conformation, and the detailed conformational analysis of tentoxin, a natural phytotoxic cyclotetrapeptide in which N-alkylated peptide bonds alternate with regular secondary ones, was used to validate the computation of conformers proportions. From the knowledge of an initial cyclic primary structure and of the D or L configuration of the amino acids, we show that it is possible to determine the exact orientation of carbonyl groups and to predict the nature of conformers present in solution. The proportion of each conformer can be inferred from a statistical thermodynamics approach by using the potential energy values of each conformer, computed by molecular mechanics methods with the TRIPOS force field, which allowed us to account for the solvent. The solvent contribution was processed by two different methods according to the nature of the interactions: whether through the dielectric constant introduced in the electrostatic potential, when interaction with solute molecules are weak or negligible, or through the computation of free energy of solvation using the algorithm SILVERWARE for solvents explicitly interacting with the solute. When applied to tentoxin, this conformational analysis yielded results in very good agreement with the experimental data reported by Pinet et al. (Biopolymers, 1995, Vol. 36, pp. 135-152), on both the nature of existing conformers and their relative proportions, whatever the nature of the considered solvent.  相似文献   

8.
A useful set of empirical rules is put forward to predict the conformations of cyclic tetrapeptides and cyclic tetradepsipeptides on the basis of primary structure, briefly presented as follows: A conformation allowing an intramolecular hydrogen bond (IMHB) of gamma-turn is preferred, and an ester bond always adopts a trans form. On a right-handed peptide ring, the carbonyl group acylating a D residue is oriented to the upper side of the main ring. The carbonyl group acylating a D proline or an N-methyl-D-amino acid residue is oriented to the lower side of the ring, forming a cis bond. The LDDL configurational sequence adopts a cis-trans-cis-trans backbone with Ci symmetry. A glycine residue behaves as a D residue in an L-peptide. Conformations of cyclotetrapeptides containing two glycine residues at diametric positions or containing an N-methyl-dehydroamino acid residue are predicted by use of appendices of rule 5. Almost all conformations of cyclic tetrapeptides are predicted by these rules. Energetical rationalization of the rules and prediction of possible new conformations are described. Conformations of cyclo(-L-Pro-L-Leu-D-Tyr(Me)-L-Ile-)(1) and cyclo (-L-Pro-D-Leu-D-Tyr(Me)-L-Ile)(2) are compared. Results of n.m.r. experiments showed that compound 1 adopts a unique cis-trans-trans-trans backbone with a gamma-turn IMHB, and 2 has a cis-trans-cis-trans backbone with Ci symmetry. These observations confirmed the rules described above. Peptides 1 and 2 are the first diastereomeric peptides with trans (LD) and cis (DD) secondary amide bonds.  相似文献   

9.
Complex formation with alkali and alkaline earth metal ions of cyclic octapeptides, cyclo(Phe-Pro)4, cyclo(Leu-Pro)4, and cyclo[Lys(Z)-Pro]4 was investigated in relation to conformation. In an alcohol solution, cyclo(Phe-Pro)4 did not form complexes. However, cyclo(Leu-Pro)4 and cyclo[Lys(Z)-Pro]4 formed complexes selectively with Ba2+ and Ca2+ ions. Changing the solvent from alcohol to acetonitrile, the complexation behavior was very different. In acetonitrile, cyclo(Phe-Pro)4 was found to form a complex with Ba2+, and CD spectra of cyclo(Leu-Pro)4 and cyclo[Lys(Z)-Pro]4 changed sharply on complexation with K+. Rate constants of the complex formation between the cyclic octapeptides and metal salts were in the range of 0.7–12 L mol?1 min?1 in an alcohol solution. One of the two types of complex formation in acetonitrile was much faster than that in an alcohol solution.  相似文献   

10.
Two new cyclic peptides, dianthins G-H (1 and 2), together with the known dianthin E (3), were isolated from the traditional Chinese medicinal plant Dianthus superbus. The sequences of cyclic peptides 1 and 2 were elucidated as cyclo (-Gly(1)-Pro(2)-Leu(3)-Thr(4)-Leu(5)-Phe(6)-) and cyclo (-Gly(1)-Pro(2)-Val(3)-Thr(4)-Ile(5)-Phe(6)-), on the basis of ESI tandem mass fragmentation analysis, extensive 2D NMR methods and X-ray diffraction. The isolated three compounds all increase proliferation of MC3T3-E1 cells in vitro using MTT method.  相似文献   

11.
In our previous study, we indicated for the first time that C-28 hydroxylation plays a significant role in the metabolism of 1alpha, 25-dihydroxyvitamin D(2) [1alpha,25(OH)(2)D(2)] by identifying 1alpha,24(S),25,28-tetrahydroxyvitamin D(2) [1alpha,24(S),25, 28(OH)(4)D(2)] as a major renal metabolite of 1alpha,25(OH)(2)D(2) [G. S. Reddy and K-Y. Tserng Biochemistry 25, 5328-5336, 1986]. The present study was performed to establish the physiological significance of C-28 hydroxylation in the metabolism of 1alpha, 25(OH)(2)D(2). We perfused rat kidneys in vitro with 1alpha, 25(OH)(2)[26,27-(3)H]D(2) (5 x 10(-10)M) and demonstrated that 1alpha,24(R),25-trihydroxyvitamin D(2) [1alpha,24(R),25(OH)(3)D(2)] and 1alpha,24(S),25,28(OH)(4)D(2) are the only two major physiological metabolites of 1alpha,25(OH)(2)D(2). In the same perfusion experiments, we also noted that there is no conversion of 1alpha,25(OH)(2)D(2) into 1alpha,25,28-trihydroxyvitamin D(2 )[1alpha,25,28(OH)(3)D(2)]. Moreover, 1alpha,24(S),25,28(OH)(4)D(2) is not formed in the perfused rat kidney when synthetic 1alpha,25, 28(OH)(3)D(2) is used as the starting substrate. This finding indicates that C-28 hydroxylation of 1alpha,25(OH)(2)D(2) occurs only after 1alpha,25(OH)(2)D(2) is hydroxylated at C-24 position. At present the enzyme responsible for the C-28 hydroxylation of 1alpha, 24(R),25(OH)(3)D(2) in rat kidney is not known. Recently, it was found that 1alpha,25(OH)(2)D(3)-24-hydroxylase (CYP24) can hydroxylate carbons 23, 24, and 26 of various vitamin D(3) compounds. Thus, it may be speculated that CYP24 may also be responsible for the C-28 hydroxylation of 1alpha,24(R),25(OH)(3)D(2) to form 1alpha, 24(S),25,28(OH)(4)D(2). The biological activity of 1alpha,24(S),25, 28(OH)(4)D(2), determined by its ability to induce intestinal calcium transport and bone calcium resorption in the rat, was found to be almost negligible. Also, 1alpha,24(S),25,28(OH)(4)D(2) exhibited very low binding affinity toward bovine thymus vitamin D receptor. These studies firmly establish that C-28 hydroxylation is an important enzymatic reaction involved in the inactivation of 1alpha,25(OH)(2)D(2) in kidney under physiological conditions.  相似文献   

12.
The RXR forms a heterodimer with the VDR to activate genes that are regulated by 1,25(OH)(2)D(3). In the absence of RXR's ligand, 9-cis-RA, RXR appears to be a silent partner to VDR. The effect of 9-cis-RA on VDR/RXR heterodimer formation and 1, 25(OH)(2)D(3)-mediated gene expression in vivo remains unclear. We examined the effect of exogenous 9-cis-RA or 9-cis-RA precursors, 9, 13-di-cis-RA and 9-cis-RCHO, on 1,25(OH)(2)D(3)-mediated induction rat renal 24-hydroxylase. The rats were treated as follows: (1) vehicle; (2) 1,25(OH)(2)D(3); (3) 1,25(OH)(2)D(3) + 9-cis-RA; (4) 1, 25(OH)(2)D(3) + 9,13-di-cis-RA; (5) 1,25(OH)(2)D(3) + 9-cis-RCHO; (6) 9-cis-RA; (7) 9,13-di-cis-RA; and (8) 9-cis-RCHO. 1, 25(OH)(2)D(3) was administered IP 18 h prior to sacrifice. The retinoids were administered every 4 h, starting 28 h prior to sacrifice. The last retinoid dose was administered 4 h prior to sacrifice. Treatment with 1,25(OH)(2)D(3) alone increased 24-hydroxylase from 35 +/- 6 (controls) to 258 +/- 44 pmol/min/g tissue. When 1,25(OH)(2)D(3) was administered with 9-cis-RA, 9, 13-di-cis-RA, or 9-cis-RCHO, 24-hydroxylases were 568 +/- 56, 524 +/- 56, and 463 +/- 62 pmol/min/g tissue, respectively. Furthermore, codosing of 1,25(OH)(2)D(3) and 9-cis-retinoids resulted in higher circulating concentrations of 9-cis-RA and 9,13-di-cis-RA when compared to rats dosed with 9-cis-retinoids alone. This was shown to be due to 1,25(OH)(2)D(3) increasing the half-life of 9,13-di-cis-RA by three to four times. These results show that 9-cis-RA can act synergistically with 1,25(OH)(2)D(3) in the regulation of 24-hydroxylase in vivo. Additionally, 1,25(OH)(2)D(3) regulates 9, 13-di-cis-RA metabolism in vivo.  相似文献   

13.
To understand trimethylamine N-oxide (TMAO) attenuation of the denaturating effects of urea or guanidine hydrochloride (GdnHCl), we have determined the apparent transfer free energies (DeltaG(tr)(')) of cyclic dipeptides (CDs) from water to TMAO, urea or GdnHCl, and also the blends of TMAO and denaturants (urea or GdnHCl) at a 1:2 ratio as well as various denaturant concentrations in the presence of 1M TMAO, through the solubility measurements, at 25 degrees C. The CDs investigated in the present study included cyclo(Gly-Gly), cyclo(Ala-Ala) and cyclo(Val-Val). The observed DeltaG(tr)(') values indicate that TMAO can stabilize the CDs while urea or GdnHCl can destabilize the CDs. Furthermore, the DeltaG(tr)(') values of the blends of TMAO with urea or GdnHCl revealed that TMAO strongly counteracted the denaturating effects of urea on CDs in all instances, however, TMAO partially counteracted the perturbing effects of GdnHCl on CDs. TMAO counteraction ability of the deleterious effects of denaturants depended on the denaturant-CDs pair. The experimental results were further used to estimate the transfer free energies (Deltag(tr)(')) of the various functional group contributions from water to TMAO, urea or GdnHCl individually and to the combinations of TMAO and the denaturants in various ratios.  相似文献   

14.
糖基化修饰是一种重要的蛋白质翻译后修饰,参与生物体中的信号传导、细胞识别等多种细胞活动,糖基缀合物的正常水解是生物体代谢的必需途径.人己糖胺酶D( Hexosaminidase D)是新发现的一种存在于人细胞质中的切除GalNAc糖基化修饰的外切酶,但该酶的酶学特性尚不清楚.利用PCR的方法,将Hex D的cDNA序列构建到质粒pET3C中,重组质粒转化大肠杆菌BL21( DE3) plysS后,通过优化异丙基-β-D-硫代吡喃半乳糖苷(IPTG)浓度(0.1mmol/L)和诱导时间(10 h)获得了高可溶性表达的重组蛋白酶.采用Ni-NTA亲和层析对重组蛋白进行了纯化,SDS-PAGE检测分子量的大小(58 kDa)和纯度(95%以上).以4-甲基伞形酮-2-乙酰氨基-2-脱氧半乳糖(4-MU-O-GalNAc)为荧光底物,测定该酶的最适反应pH值为5.5,最适反应温度为37℃,且该酶的热稳定性较好,在50℃下放置半小时仍有较高活性,1mmol/L的金属离子(CuSO4、FeSO4·7H2O、MgCl2· 6H2O、CaCl2、NiSO4·6H2O、AlCl3·6H2O、ZnSO4·7H2O、MnCl2)及EDTA对该酶活性影响不大,10mmol/L AlCl3、CuSO、FeSO4·7H2O对该酶有不同程度的抑制.在最适条件下(pH 5.5,37℃)下,该酶的Km为0.16mmoL/L,最大反应速率为3.06 μmol/( min·mg).  相似文献   

15.
cyclo(-L-Pro-L-Val-L-Pro-L-Val-) (1L) and cyclo(-L-Pro-D-Val-L-Pro-D-Val-) (1D) were synthesized by the conventional method for peptide synthesis. Conformations of 1L and 1D in solution were studied. Compound 1L has a cis-trans-cis-trans backbone conformation with C2 symmetry in CDCl3. This conformation is slightly different from that in crystalline state and in DMSO-d6 solution. Compound 1D has a cis-trans-cis-trans conformation in DMSO-d6 and an all-trans conformation in trifluoroethanol-d3. Compound 1L retarded stem-growth of rice seedlings and, in contrast, compound 1D promoted root-growth of rice seedlings.  相似文献   

16.
The active vitamin D(3)-metabolite 1,25(OH)(2)D(3) inhibits the interleukin 4/granulocyte-macrophage colony-stimulating factor (IL-4/GM-CSF)-induced differentiation of human monocytes into dendritic cells without altering survival. Colony-stimulating factor 1 (CSF-1) is an important survival factor for cells of the monocytic lineage. We therefore investigated whether the inhibitory activity of 1,25(OH)(2)D(3) is paralleled by a regulation of CSF-1 and its receptor. Purified human monocytes were cultured together with IL-4/GM-CSF in the presence of 1,25(OH)(2)D(3), its analogue tacalcitol, the low-affinity vitamin D receptor ligand 24,25(OH)(2)D(3), or the solvent ethanol for up to 5 days. Expression of CSF-1, CSF-1R, and GM-CSF mRNA was measured by RT-PCR. Protein secretion for CSF-1 was measured by ELISA, expression of CSF-1R by flow cytometry. The results showed that 1,25(OH)(2)D(3) and tacalcitol significantly up-regulated CSF-1 mRNA-expression and protein secretion in a dose-dependent manner. The effect of 1,25(OH)(2)D(3) occurred already after 1h of pre-treatment. In contrast, CSF-1R mRNA- and cell surface-expression was down-regulated simultaneously. The solvent ethanol and 24,25(OH)(2)D(3) were without effect. GM-CSF mRNA expression was not modulated in 1,25(OH)(2)D(3)-treated cells. These data point towards a distinct and specific regulation of CSF-1 and its receptor by 1,25(OH)(2)D(3) and its analogue tacalcitol in human monocytes which parallels the inhibition of differentiation into dendritic cells without altering survival.  相似文献   

17.
H N Bhargava  P Ramarao 《Life sciences》1989,45(26):2521-2528
The effect of cyclo(Leu-Gly) on U-50,488H- induced pharmacological actions was determined in male Sprague-Dawley rats. Intraperitoneal (i.p.) administration of U-50,488H to rats produced analgesia (tail-flick) and increased urinary output. Cyclo (Leu-Gly) (1-4 mg/kg, s.c.) antagonized the analgesic response to U-50,488H (25 mg/kg; i.p.). A dose of 10 mg/kg (i.p.) of U-50,488H increased the spontaneous urinary output which was antagonized by cyclo (Leu-Gly) (1-4 mg/kg; s.c.). To determine whether cyclo (Leu-Gly) was acting as a kappa-opioid receptor antagonist, the effect of cyclo (Leu-Gly) on the binding of [3H]ethylketocyclazocine (EKC) to membranes of rat cerebral cortex and spinal cord was determined. The IC50 values of cyclo(Leu-Gly) in displacing [3H]EKC from its binding sites in cortex and spinal cord were 1.44 and 0.40 mM, respectively. Chronic administration of U-50,488H (25 mg/kg; i.p., b.i.d.) for 4 days induced tolerance to its analgesic effect. The latter was not affected by cyclo(Leu-Gly) (2 to 8 mg/kg; s.c.) given once a day for 4 days. It is concluded that cyclo(Leu-Gly) antagonizes acute actions of U-50,488H and that such effects of cyclo(Leu-Gly) are not mediated via a direct action on kappa-opioid receptors.  相似文献   

18.
This study is aimed at elucidating the structure of a novel T‐cell adhesion inhibitor, cyclo(1,8)‐CPRGGSVC using one‐ and two‐dimensional (2D) 1H NMR and molecular dynamics (MD) simulation. The peptide is derived from the sequence of its parent peptide cIBR (cyclo(1,12)‐PenPRGGSVLVTGC), which is a fragment of intercellular adhesion molecule‐1 (ICAM‐1). Our previous results show that the cyclo(1,8)‐CPRGGSVC peptide binds to the LFA‐1 I‐domain and inhibits heterotypic T‐cell adhesion, presumably by blocking the LFA‐1/ICAM‐1 interactions. The structure of the peptide was determined using NMR and MD simulation in aqueous solution. Our results indicate that the peptide adopts type‐I β‐turn conformation at the Pro2‐Arg3‐Gly4‐Gly5 (PRGG) sequence. The β‐turn structure at the PRGG motif is well conserved in cIBR peptide and ICAM‐1 receptor, which suggests the importance of the PRGG motif for the biological activity of cyclo(1,8)‐CPRGGSVC peptide. Meanwhile, the Gly5‐Ser6‐Val7‐Cys8‐Cys1 (GSVCC) sequence forms a “turn‐like” random coil structure that does not belong to any structured motif. Therefore, cyclo(1,8)‐CPRGGSVC peptide has only one structured region at the PRGG sequence, which may play an important role in the binding of the peptide to the LFA‐1 I‐domain. The conserved β‐turn conformation of the PRGG motif in ICAM‐1, cIBR, and cyclo(1,8)‐CPRGGSVC peptides can potentially be used to design peptidomimetics. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 633–641, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

19.
Amphiphilic ligands 4′-((4-(5-pyridin-4-yloxy)pentyloxy)styryl)-4-methyl-2,2′-bipyridine (L1), 4,4′-bis((E)-4-(5-(pyridin-4-yloxy)pentyloxy)styryl)-2,2′-bipyridine (L2), 4′-(4-(5-(4′-cyano-4-biphenoxy)pentyloxy)styryl)-4-methyl-2,2′-bipyridine (L3), and 4′-(4-(5-(zinc tetrakis-5,10,15-tritolyl-20-(4-hydroxyphenyl)porphyrin)pentyloxy)styryl)-4-methyl-2,2′-bipyridine (L4) and their heteroleptic ruthenium(II) complexes of the types [Ru(L1)(L)(NCS)2] (D20), [Ru(L2)(L)(NCS)2] (D21), [Ru(L3)(L)(NCS)2] (D22), and [Ru(L4)(L)(NCS)2] (D23) (where L = 4,4′-bis(carboxylic acid)-2,2′-bipyridine) have been synthesized, as photosensitizers for nanocrystalline dye-sensitized solar cells. All complexes D20-D23 exhibit a broad MLCT band around 520-530 nm in DMF and an emission band around 740-790 nm in DMF. We have studied photovoltaic performances based on the newly synthesized dyes. Under standard AM 1.5 sunlight, the dye D20 gave a short-circuit photocurrent density of 13.31 mA/cm2, an open-circuit voltage of 0.64 V, and a fill factor of 0.68, corresponding to an overall conversion efficiency of 5.81%.  相似文献   

20.
It has been proposed that the membrane allows a much more efficient binding of certain small or medium-sized amphiphilic messenger molecules to their receptor, not only by accumulation of the drug, but also by induction of orientations and conformations that are much more favorable for receptor docking than structures adopted in isotropic phases. A series of eight amphiphilic cyclic peptides containing lipophilic (L-alpha-aminodecanoic acid = Ada, L-alpha-aminohexadecanoic acid = Ahd, Nhdg = N-hexadecylglycine) and hydrophilic (Lys, Asp) amino acids were synthesized and examined by means of NMR spectroscopy and molecular dynamics (MD) simulations in isotropic (CDCl3) and membrane-mimicking anisotropic (SDS/H2O) solvents to study the influence of the environment on their individual conformations. NMR data of cyclo(-Gly1-D-Asp2-Ahd3-Ahd4-Asp5-Gly6+ ++-) (C4), cyclo(-Lys1-D-Pro2-Lys3-Ada4-Pro5-Ada6-) (C5) and cyclo(-Lys1-Pro2-Lys3-Ada4-D-Pro5-Ada6-) (C6) clearly indicate that those compounds are too rigid to perform a conformational change upon transition from an isotropic to an anisotropic environment. On the other hand, the experimental data of cyclo (-Gly1-Asp2-Ahd3-Ahd4-Asp5-Gly6-) (C1), cyclo(-Asp1-Ala2-Nhdg3-Ala4-D-Asp5-) (C7), and cyclo(-D-Asp1-Ala2-Nhdg3-Ala4-Asp5-) (C8) suggest highly flexible unstructured molecules in both environments. However, for cyclo(-Asp1-Asp2-Gly3-Ahd4-Ahd5-Gly6-) (C2) we observed a structure inducing effect of a membrane-like environment. The compound populates three different conformations in SDS/H2O, whereas in CDCI3 no preferred conformation can be detected. cyclo(-D-Asp1-Asp2-Gly3-Ahd4-Ahd5-Gly6-) (C3) clearly exhibits two different conformations with a shifted beta,beta-turn motif in CDCI3 and SDS/H2O solutions. The conformational change could be reproduced in a restraint-free MD simulation using the biphasic membrane mimetic CCl4/H2O. Our results give clear evidence that membrane interactions may not only lead to structure inductions, but can also induce major conformational changes in compounds already exhibiting a defined structure in isotropic solution.  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