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.
Poly‐ε‐lysine produced by streptomyces species is a promising biopolymer owing to its antimicrobial activity and safety for humans. A number of nutritional factors influencing poly‐ε‐lysine production by Streptomyces noursei NRRL 5126 were studied. Various metabolic precursors such as amino acids, tricarboxylic acid cycle intermediates and cofactors were investigated for improved production of poly‐ε‐lysine. Results indicated L ‐aspartate (2 mM) and citric acid (5 mM) to substantially increase the poly‐ε‐lysine production from 97.08 to 409.94 mg/L. Addition of citric acid after 24 h and L ‐aspartate after 36 h of fermentation medium further enhanced poly‐ε‐lysine production to 497.67 mg/L after a total fermentation time of 108 h. However, the use of cofactors of enzymes involved in the biosynthesis of poly‐ε‐lysine inhibited its production which is believed to be due to diversion of the flux to other metabolites.  相似文献   

2.
L ‐Lysine is a potential feedstock for the production of bio‐based precursors for engineering plastics. In this study, we developed a microbial process for high‐level conversion of L ‐lysine into 5‐aminovalerate (5AVA) that can be used as a monomer in nylon 6,5 synthesis. Recombinant Escherichia coli WL3110 strain expressing Pseudomonas putida delta‐aminovaleramidase (DavA) and lysine 2‐monooxygenase (DavB) was grown to high density in fed‐batch culture and used as a whole cell catalyst. High‐density E. coli WL3110 expressing DavAB, grown to an optical density at 600 nm (OD600) of 30, yielded 36.51 g/L 5AVA from 60 g/L L ‐lysine in 24 h. Doubling the cell density of E. coli WL3110 improved the conversion yield to 47.96 g/L 5AVA from 60 g/L of L ‐lysine in 24 h. 5AVA production was further improved by doubling the L ‐lysine concentration from 60 to 120 g/L. The highest 5AVA titer (90.59 g/L; molar yield 0.942) was obtained from 120 g/L L ‐lysine by E. coli WL3110 cells grown to OD600 of 60. Finally, nylon 6,5 was synthesized by bulk polymerization of ?‐caprolactam and δ‐valerolactam prepared from microbially synthesized 5AVA. The hybrid system demonstrated here has promising possibilities for application in the development of industrial bio‐nylon production processes.  相似文献   

3.
L ‐Homophenylalanine (L ‐HPA) and N6‐protected‐2‐oxo‐6‐amino‐hexanoic acid (N6‐protected‐OAHA) can be used as building blocks for the manufacture of angiotensin‐converting enzyme inhibitors. To synthesize L ‐HPA and N6‐protected‐OAHA simultaneously from 2‐oxo‐4‐phenylbutanoic acid (OPBA) and N6‐protected‐L ‐lysine, several variants of Escherichia coli aspartate aminotransferase (AAT) were developed by site‐directed mutagenesis and their catalytic activities were investigated. Three kinds of N6‐protected‐L ‐lysine were tested as potential amino donors for the bioconversion process. AAT variants of R292E/L18H and R292E/L18T exhibited specific activities of 0.70±0.01 U/mg protein and 0.67±0.02 U/mg protein to 2‐amino‐6‐tert‐butoxycarbonylamino‐hexanoic acid (BOC‐lysine) and 2‐amino‐6‐(2,2,2‐trifluoro‐acetylamino)‐hexanoic acid, respectively. E. coli cells expressing R292E/L18H variant were able to convert OPBA and BOC‐lysine to L ‐HPA and 2‐oxo‐6‐tert‐butoxycarbonylamino‐hexanoic acid (BOC‐OAHA) with 96.2% yield in 8 h. This is the first report demonstrating a process for the simultaneous production of two useful building blocks, L ‐HPA and BOC‐OAHA. © 2009 American Institute of Chemical Engineers Biotechnol. Prog., 2009  相似文献   

4.
Synchrotron ir radiation was used to collect far‐ir spectroscopy data on amino acid homopolymers, including poly‐L ‐phenylalanine, poly‐L ‐alanine, poly‐L ‐leucine, poly‐L ‐valine, and poly‐L ‐tryptophan. The intensity of the broadband synchrotron ir source, coupled with the use of a Michelson interferometer and Fourier transformed data, provides uniquely high signal to noise far‐ir data, which can be analyzed by deconvolution techniques that are well established. The temperature dependence of the spectra reveals consistent trends, with the bands shifting to higher frequencies and intensities as the temperature is lowered. A careful analysis of poly‐L ‐phenylalanine, as a function of temperature from 10 to 295 K, reveals consistent (monotonic) changes with temperature. An explanation for these effects is presented. Since the observed modes are of low frequency and low energy, a substantial fraction of the observed transitions result from a combination of excited state absorption and stimulated emission at ambient temperature. Anharmonicity of the vibrational potentials, with the resultant decrease in spacing of higher energy levels, is consistent with the frequency up shifts as temperature is decreased. The results emphasize that these excited state effects likely govern much of the behavior of low‐frequency modes in all biomolecules. © 1999 John Wiley & Sons, Inc. Biopoly 49: 591–603, 1999  相似文献   

5.
6.
Cardiac fibroblasts (CFs) regulate myocardial remodeling by proliferating, differentiating, and secreting extracellular matrix (ECM) proteins. B‐type natriuretic peptide (BNP) is anti‐fibrotic, inhibits collagen production, augments matrix metalloproteinases, and suppresses CF proliferation. Recently, we demonstrated that the ECM protein fibronectin (FN) augmented production of BNP's second messenger, 3′, 5′ cyclic guanosine monophosphate (cGMP) in CFs, supporting crosstalk between FN, BNP, and its receptor, natriuretic peptide receptor A (NPR‐A). Here, we address the specificity of FN to augment cGMP generation by investigating other matrix proteins, including collagen IV which contains RGD motifs and collagen I and poly‐L ‐lysine, which have no RGD domain. Collagen IV showed increased cGMP generation to BNP similar to FN. Collagen I and poly‐L ‐lysine had no effect. As FN also interacts with integrins, we then examined the effect of integrin receptor antibody blockade on BNP‐mediated cGMP production. On FN plates, antibodies blocking RGD‐binding domains of several integrin subtypes had little effect, while a non‐RGD domain interfering integrin αvβ3 antibody augmented cGMP production. Further, on uncoated plates, integrin αvβ3 blockade continued to potentiate the BNP/cGMP response. These studies suggest that both RGD containing ECM proteins and integrins may interact with BNP/NPR‐A to modulate cGMP generation. J. Cell. Physiol. 225: 251–255, 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

7.
Four diastereomeric‐Leu‐Leu‐Aib‐Leu‐Leu‐Aib‐peptides, Boc‐D ‐Leu‐L ‐Leu‐Aib‐L ‐Leu‐L ‐Leu‐Aib‐OMe (1), Boc‐L ‐Leu‐D ‐Leu‐Aib‐L ‐Leu‐L ‐Leu‐Aib‐OMe (2), Boc‐L ‐Leu‐L ‐Leu‐Aib‐D ‐Leu‐L ‐Leu‐Aib‐OMe (3), and Boc‐L ‐Leu‐L ‐Leu‐Aib‐L ‐Leu‐D ‐Leu‐Aib‐OMe (4), were synthesized. The crystals of the four hexapeptides were characterized by X‐ray crystallographic analysis. Two diastereomeric hexapeptides 1 and 2 having D ‐Leu(1) or D ‐Leu(2) were folded into right‐handed (P) 3 10 ‐helical structures, while peptide 3 having D ‐Leu(4) was folded into a turn structure nucleated by type III′ and I$' \bf{\beta}$ ‐turns, and peptide 4 having D ‐Leu(5) was folded into a left‐handed (M) 3 10 ‐helical structure. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

8.
The interactions of poly‐L ‐glutamic acid and a cationic porphyrin derivative in aqueous solutions were studied by the combination of vibrational circular dichroism (VCD) and electronic circular dichroism (ECD) spectroscopies. It was found that non‐covalent interactions between both agents influence the structure of the polymeric matrix and the guest porphyrins and vice versa, but the physico‐chemical properties of the solutions, especially the pH and the relative permittivity of the solvent, play a key role in the structure of the polypeptide part of the formed complexes. It was shown that the interaction with porphyrins prevents the precipitation of poly‐L ‐glutamic acid in aqueous solution at acidic pH. In special conditions, the porphyrins attached to the polypeptide probably possess face‐to‐face interaction as demonstrated by the enhancement of the characteristic ECD signal and the appearance of sidebands on its short and long wavelength sides. Copyright © 2005 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

9.
Under the hypotheses of a structurally related binding site for antagonists of G‐protein coupled receptors and the ability of cyclic pentapeptides of chiral sequence D 1L 2D 3D 4L 5 to form rigid structures with which probe the pharmacophoric specificity of these receptors, inhibitors of substance P were designed based on available structure–activity relationships. ITF 1565, cyclo[D ‐Trp1‐Pro2‐D ‐Lys3‐D ‐Trp4‐Phe5], antagonized substance P activity mediated by type 1 neurokinin receptor (NK1) whereas it acted weakly against NK2 and did not inhibit endothelin at all. The preferential conformation of the peptide was obtained from nmr spectroscopy and computer calculations, and shown to contain the same βII‐turn and γ′‐turn found in other cyclic pentapeptides with the same chiral sequence. The structure of the peptide was compared with that of the β‐D ‐glucose molecule that has been proposed as a semirigid scaffold for antagonists of G‐protein coupled receptors. The γ′‐turn of the cyclic peptide superimposed well with β‐D ‐glucose in the chair conformation. Furthermore, when the side chains were considered, the aromatic groups of the two molecules were found to generally overlap. These results support the view of G‐protein coupled receptors as possessing structurally similar binding sites for antagonists and suggest that cyclic pentapeptides of chiral sequence D 1L 2D 3D 4L 5 may be useful as semirigid scaffolds for the design of antagonists of this family of receptors. © 1999 John Wiley & Sons, Inc. Biopoly 50: 211–219, 1999  相似文献   

10.
Diffusion into and from κ‐carrageenan gel beads was studied, both in the absence and presence of bacterial cells, both with and without biochemical reaction. The solutes were indole, L ‐serine, and L ‐tryptophan. The reaction was that of indole and L ‐serine to give L ‐tryptophan. Established theory concerning diffusion of a single solute in cell‐free gels was found to describe well the effect of the gel on diffusivity. Simultaneous diffusion of the three solutes resulted in lower diffusivities than those for individual solutes, suggesting the need to use multicomponent diffusion theory. The effect of cells on diffusion could only be accounted for by models assuming permeable cells. Diffusion with chemical reaction was reasonably well described by an effectiveness factor calculated using an effective diffusivity estimated from diffusion data without reaction. © 1999 John Wiley & Sons, Inc. Biotechnol Bioeng 63: 625–631, 1999.  相似文献   

11.
Sperm‐associated α‐L ‐fucosidases have been implicated in fertilization in many species. Previously, we documented the existence of α‐L ‐fucosidase in mouse cauda epididymal contents, and showed that sperm‐associated α‐L ‐fucosidase is cryptically stored within the acrosome and reappears within the sperm equatorial segment after the acrosome reaction. The enrichment of sperm membrane‐associated α‐L ‐fucosidase within the equatorial segment of acrosome‐reacted cells implicates its roles during fertilization. Here, we document the absence of α‐L ‐fucosidase in mouse oocytes and early embryos, and define roles of sperm associated α‐L ‐fucosidase in fertilization using specific inhibitors and competitors. Mouse sperm were pretreated with deoxyfuconojirimycin (DFJ, an inhibitor of α‐L ‐fucosidase) or with anti‐fucosidase antibody; alternatively, mouse oocytes were pretreated with purified human liver α‐L ‐fucosidase. Five‐millimolar DFJ did not inhibit sperm–zona pellucida (ZP) binding, membrane binding, or fusion and penetration, but anti‐fucosidase antibody and purified human liver α‐L ‐fucosidase significantly decreased the frequency of these events. To evaluate sperm‐associated α‐L ‐fucosidase enzyme activity in post‐fusion events, DFJ‐pretreated sperm were microinjected into oocytes, and 2‐pronuclear (2‐PN) embryos were treated with 5 mM DFJ with no significant effects, suggesting that α‐L ‐fucosidase enzyme activity does not play a role in post‐fusion events and/or early embryo development in mice. The recognition and binding of mouse sperm to the ZP and oolemma involves the glycoprotein structure of α‐L ‐fucosidase, but not its catalytic action. These observations suggest that deficits in fucosidase protein and/or the presence of anti‐fucosidase antibody may be responsible for some types of infertility. Mol. Reprod. Dev. 80: 273–285, 2013. © 2013 Wiley Periodicals, Inc.  相似文献   

12.
We studied the interaction of poly‐l ‐lysine (PLL) and poly‐l ‐arginine (PLAG) with sodium dodecyl sulfate (SDS) surfactant and the interaction of poly‐l‐ glutamic acid (PLGA) and poly‐l ‐aspartic acid (PLAA) with tetradecyltrimethylammonium bromide (TTAB) surfactant using vibrational circular dichroism (VCD) spectroscopy in the region of C‐H stretching vibration and in the Amide I region both in solution and in mulls. A chirality transfer from polypeptides to achiral surfactants was observed in the C‐H stretching region, where measurements in solution were impossible. This observation was enabled by a special sample treatment technique using lyophilization and the preparation of mulls. This technique demonstrated itself as an interesting and beneficial tool for VCD measurements. In addition, we observed that SDS changed the secondary structure of PLL to the β‐sheet and of PLAG to the α‐helix. TTAB disrupted the PLGA and PLAA structure. These results were obtained in the mull but were confirmed by the VCD spectra measured in solution and by electronic circular dichroism. The chirality transfer from the polypeptides to SDS was caused by polypeptides ordered into a specific conformation during the interaction, while in the TTBA system it was induced primarily by the chirality of the amino acid residues. Chirality 27:965–972, 2015. © 2015 Wiley Periodicals, Inc.  相似文献   

13.
Studies of the helix‐to‐coil transition in dilute solutions of poly‐L ‐lysine, dissolved in mixtures of water and methanol (MeOH), have been carried under shear flow using flow birefringence and modulated polarimetry. The fraction of helical conformations in a given solution remains independent of shear rate for MeOH concentrations above and below the critical value for the helix‐coil transition (i.e., 87.5% MeOH). For the 87.5% MeOH solutions, a shear‐induced helix‐to‐“stretched” coil transition occurs above a critical shear rate. Induction times for the transition show a temperature and shear rate dependence that can be described in terms of an activated jump process. Measurements of circular birefringence on cessation of flow also show that the transition is reversible, with the stretched coil reverting to the helical state on a time scale of several seconds. The activation energy for the jump process is found to be 16.2 kJ/mole. © 1999 John Wiley & Sons, Inc. Biopoly 50: 589–594, 1999  相似文献   

14.
In the brain, the human flavoprotein D ‐amino acid oxidase (hDAAO) is involved in the degradation of the gliotransmitter D ‐serine, an important modulator of NMDA‐receptor‐mediated neurotransmission; an increase in hDAAO activity (that yields a decrease in D ‐serine concentration) was recently proposed to be among the molecular mechanisms leading to the onset of schizophrenia susceptibility. This human flavoenzyme is a stable homodimer (even in the apoprotein form) that distinguishes from known D ‐amino acid oxidases because it shows the weakest interaction with the flavin cofactor in the free form. Instead, cofactor binding is significantly tighter in the presence of an active site ligand. In order to understand how hDAAO activity is modulated, we investigated the FAD binding process to the apoprotein moiety and compared the folding and stability properties of the holoenzyme and the apoprotein forms. The apoprotein of hDAAO can be distinguished from the holoenzyme form by the more “open” tertiary structure, higher protein fluorescence, larger exposure of hydrophobic surfaces, and higher sensitivity to proteolysis. Interestingly, the FAD binding only slightly increases the stability of hDAAO to denaturation by urea or temperature. Taken together, these results indicate that the weak cofactor binding is not related to protein (de)stabilization or oligomerization (as instead observed for the homologous enzyme from yeast) but rather should represent a means of modulating the activity of hDAAO. We propose that the absence in vivo of an active site ligand/substrate weakens the cofactor binding, yielding the inactive apoprotein form and thus avoiding excessive D ‐serine degradation.  相似文献   

15.
Insulin causes endothelium‐derived nitric oxide (NO)‐dependent vascular relaxation, and increases L ‐arginine transport via cationic amino acid transporter 1 (hCAT‐1) and endothelial NO synthase (eNOS) expression and activity in human umbilical vein endothelium (HUVEC). We studied insulin effect on SLC7A1 gene (hCAT‐1) expression and hCAT‐transport activity role in insulin‐modulated human fetal vascular reactivity. HUVEC were used for L ‐arginine transport and L ‐[3H]citrulline formation (NOS activity) assays in absence or presence of N‐ethylmaleimide (NEM) or L ‐lysine (L ‐arginine transport inhibitors). hCAT‐1 protein abundance was estimated by Western blot, mRNA quantification by real time PCR, and SLC7A1 promoter activity by Luciferase activity (?1,606 and ?650 bp promoter fragments from ATG). Specific protein 1 (Sp1), and total or phosphorylated eNOS protein was determined by Western blot. Sp1 activity (at four sites between ?177 and ?105 bp from ATG) was assayed by chromatin immunoprecipitation (ChIP) and vascular reactivity in umbilical vein rings. Insulin increased hCATs–L ‐arginine transport, maximal transport capacity (Vmax/Km), and hCAT‐1 expression. NEM and L ‐lysine blocked L ‐arginine transport. In addition, it was trans‐stimulated (~7.8‐fold) by L ‐lysine in absence of insulin, but unaltered (~1.4‐fold) in presence of insulin. Sp1 nuclear protein abundance and binding to DNA, and SLC7A1 promoter activity was increased by insulin. Insulin increased NO synthesis and caused endothelium‐dependent vessel relaxation and reduced U46619‐induced contraction, effects blocked by NEM and L ‐lysine, and dependent on extracellular L ‐arginine. We suggest that insulin induces human umbilical vein relaxation by increasing HUVEC L ‐arginine transport via hCATs (likely hCAT‐1) most likely requiring Sp1‐activated SLC7A1 expression. J. Cell. Physiol. 226: 2916–2924, 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

16.
The mannosylated derivative of adamant‐1‐yl tripeptide (D ‐(Ad‐1‐yl)Gly‐L ‐Ala‐D ‐isoGln) was prepared to study the effects of mannosylation on adjuvant (immunostimulating) activity. Mannosylated adamant‐1‐yl tripeptide (Man‐OCH2CH(Me)CO‐D ‐(Ad‐1‐yl)Gly‐L ‐Ala‐D ‐isoGln) is a non‐pyrogenic, H2O‐soluble, and non‐toxic compound. Adjuvant activity of mannosylated adamantyl tripeptide was tested in the mouse model with ovalbumin as an antigen and in comparison to the parent tripeptide and peptidoglycan monomer (PGM, β‐D ‐GlcNAc‐(1→4)‐D ‐MurNAc‐L ‐Ala‐D ‐isoGln‐mesoDAP(εNH2)‐D ‐Ala‐D ‐Ala), a well‐known effective adjuvant. The mannosylation of adamantyl tripeptide caused the amplification of its immunostimulating activity in such a way that it was comparable to that of PGM.  相似文献   

17.
Both the molecular recognition and interaction of metallo‐β‐lactamase CcrA with l ‐captopril were studied by the combined use of fluorescence spectra and molecular dynamic simulation. The results showed that the binding constant was 8.89 × 104 L mol?1 at 296 K. Both Zn1 and Zn2 displayed tetrahedral coordination geometries in the CcrA–Lcap complex, the S atom in l ‐captopril displaced the nucleophilic hydroxide in apo CcrA and occupied the fourth coordination site for each ion, resulting in a competitively inhibited CcrA enzyme. Strong electrostatic interaction between the two zinc ions in CcrA and negatively charged l ‐captopril provided the main driving force for the binding affinity. Through a partly structural transformation from β‐sheet to random coil, loop 1 (residues 24–34) completely opened the binding pocket of CcrA to allow an induced fit of the newly introduced ligand. This study may provide some valuable information for designing and developing a more tightly binding inhibitor to resist superbugs.  相似文献   

18.
The aim of this work was to prepare L ‐ and D ‐(adamant‐1‐yl)‐Gly‐L ‐Ala‐D ‐isoGln peptides in order to study their adjuvant (immunostimulating) activities. Adjuvant activity of adamant‐1‐yl tripeptides was tested in the mouse model using ovalbumin as an antigen and in comparison to the peptidoglycan monomer (PGM; β‐D ‐GlcNAc‐(1→4)‐D ‐MurNAc‐L ‐Ala‐D ‐isoGln‐mesoDAP(εNH2)‐D ‐Ala‐D ‐Ala) and structurally related adamant‐2‐yl tripeptides.  相似文献   

19.
The interactions of lysine oligopeptides with dimyristoyl phosphatidylglycerol (DMPG) bilayer membranes were studied using spin-labeled lipids and electron spin resonance spectroscopy. Tetralysine and pentalysine were chosen as models for the basic amino acid clusters found in a variety of cytoplasmic membrane-associating proteins, and polylysine was chosen as representative of highly basic peripherally bound proteins. A greater motional restriction of the lipid chains was found with increasing length of the peptide, while the saturation ratio of lipids per peptide was lower for the shorter peptides. In DMPG and dimyristoylphosphatidylserine host membranes, the perturbation of the lipid chain mobility by polylysine was greater for negatively charged spin-labeled lipids than for zwitterionic lipids, but for the shorter lysine peptides these differences were smaller. In mixed bilayers composed of DMPG and dimyristoylphosphatidylcholine, little difference was found in selectivity between spin-labeled phospholipid species on binding pentalysine. Surface binding of the basic lysine peptides strongly reduced the interfacial pK of spin-labeled fatty acid incorporated into the DMPG bilayers, to a greater extent for polylysine than for tetralysine or pentalysine at saturation. The results are consistent with a predominantly electrostatic interaction with the shorter lysine peptides, but with a closer surface association with the longer polylysine peptide.  相似文献   

20.
We report the conformational analysis by 1H‐nmr and computer simulations of five potent sweet molecules, N‐(3,3‐dimethylbutyl)‐L ‐aspartyl‐S‐(α‐methyl)phenylalanine methylester (1; 5000 times more potent than sucrose), L ‐aspartyl‐D ‐valine (S)‐α‐methoxycarbonylmethylbenzylamide (2; 1400 times more potent than sucrose), L ‐aspartyl‐D ‐valine α‐phenylcyclopentylamide (3; 1200 times more potent than sucrose), L ‐aspartyl‐D ‐α‐aminobutyric acid (S)‐α‐cyclohexylpropylamide (4; 2300 times more potent than sucrose), and L ‐aspartyl‐D ‐valine (R)‐α‐methylthiomethylbenzylamide (5; 3000 times more potent than sucrose). The “L‐shaped” structure, which we believe to be responsible for sweet taste, is accessible to all five sweet compounds in solution. This structure is characterized by a zwitterionic ring formed by the A‐H and B containing moieties located in the +y axis and by the hydrophobic group X pointing into the +x axis. Other accessible conformations of these flexible molecules are extended conformations with the A‐H and B containing moieties in the +y axis and the hydrophobic group X pointing in the −y axis and reversed L‐shaped structures with the hydrophobic group X projecting along the −x axis. The remarkable potency of the N‐alkylated compound 1 supports our recent hypothesis that a second hydrophobic binding domain in addition to interactions arising from the L‐shaped structure leads to an enhancement of sweetness potency. © 1999 John Wiley & Sons, Inc. Biopoly 49: 525–539, 1999  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