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.
2.
The kinetics of interaction of Esigma(70) RNA polymerase (R) with the lambdaP(R) promoter (P) were investigated by filter binding over a broad range of temperatures (7.3-42 degrees C) and concentrations of RNA polymerase (1-123 nM) in large excess over promoter DNA. Under all conditions examined, the kinetics of formation of competitor-resistant complexes (I(2), RP(o)) are single-exponential with first order rate constant beta(CR). Interpretation of the polymerase concentration dependence of beta(CR) in terms of the three step mechanism of open complex formation yields the equilibrium constant K(1) for formation of the first kinetically significant intermediate (I(1)) and the forward rate constant (k(2)) for the conformational change converting I(1) to the second kinetically significant intermediate I(2): R + P-->(K(1))<--I(1)(k(2))-->I(2). Use of rapid quench mixing allows K(1) and k(2) to be individually determined over the entire temperature range investigated, previously not possible at this promoter using manual mixing. Given the large (>60 bp) interface formed in I(1), its relatively small binding constant K(1) at 37 degrees C at this [salt] (approximately 6 x 10(6) M(-1)) strongly argues that binding free energy is used to drive large-scale structural changes in polymerase and/or promoter DNA or other coupled processes. Evidence for coupling of protein folding is provided by the large and negative activation heat capacity of k(a)[DeltaC(o,++)(a)= -1.5(+/-0.2)kcal K(-1)], now shown to originate directly from formation of I(1) [DeltaC(o)(1)= -1.4(+/-0.3)kcal K(-1)] rather than from the formation of I(2) as previously proposed. The isomerization I(1)-->I(2) exhibits relatively slow kinetics and has a very large temperature-independent Arrhenius activation energy [E(act)(2)= 34(+/-2)kcal]. This kinetic signature suggests that formation of the transition state (I(1)-I(2)++ involves large conformational changes dominated by changes in the exposure of polar and/or charged surface to water. Structural and biochemical data lead to the following hypotheses to interpret these results. We propose that formation of I(1) involves coupled folding of unstructured regions of polymerase (beta, beta' and sigma(70)) and bending of promoter DNA (in the -10 region). We propose that interactions with region 2 of sigma(70) and possibly domain 1 of beta induce a kink at the -11/-12 base pairs of the lambdaP(R) promoter which places the downstream DNA (-5 to +20) in the jaws of the beta and beta' subunits of polymerase in I(1). These early interactions of beta and beta' with the DNA downstream of position -5 trigger jaw closing (with coupled folding) and subsequent steps of DNA opening.  相似文献   

3.
4.
During iron acquisition by the cell, complete homodimeric transferrin receptor 1 in an unknown state (R1) binds iron-loaded human serum apotransferrin in an unknown state (T) and allows its internalization in the cytoplasm. T also forms complexes with metals other than iron. Are these metals incorporated by the iron acquisition pathway and how can other proteins interact with R1? We report here a four-step mechanism for cobalt(III) transfer from CoNtaCO(3)(2-) to T and analyze the interaction of cobalt-loaded transferrin with R1. The first step in cobalt uptake by T is a fast transfer of Co(3+) and CO(3)(2-) from CoNtaCO(3)(2-) to the metal-binding site in the C-lobe of T: direct rate constant, k(1)=(1.1+/-0.1) x 10(6) M(-1) s(-1); reverse rate constant, k(-1)=(1.9+/-0.6) x 10(6) M(-1) s(-1); and equilibrium constant, K=1.7+/-0.7. This step is followed by a proton-assisted conformational change of the C-lobe: direct rate constant, k(2)=(3+/-0.3) x 10(6) M(-1) s(-1); reverse rate constant, k(-2)=(1.6+/-0.3) x 10(-2) s(-1); and equilibrium constant, K(2a)=5.3+/-1.5 nM. The two final steps are slow changes in the conformation of the protein (0.5 h and 72 h), which allow it to achieve its final thermodynamic state and also to acquire second cobalt. The cobalt-saturated transferrin in an unknown state (TCo(2)) interacts with R1 in two different steps. The first is an ultra-fast interaction of the C-lobe of TCo(2) with the helical domain of R1: direct rate constant, k(3)=(4.4+/-0.6)x10(10) M(-1) s(-1); reverse rate constant, k(-3)=(3.6+/-0.6) x 10(4) s(-1); and dissociation constant, K(1d)=0.82+/-0.25 muM. The second is a very slow interaction of the N-lobe of TCo(2) with the protease-like domain of R1. This increases the stability of the protein-protein adduct by 30-fold with an average overall dissociation constant K(d)=25+/-10 nM. The main trigger in the R1-mediated iron acquisition is the ultra-fast interaction of the metal-loaded C-lobe of T with R1. This step is much faster than endocytosis, which in turn is much faster than the interaction of the N-lobe of T with the protease-like domain. This can explain why other metal-loaded transferrins or a protein such as HFE-with a lower affinity for R1 than iron-saturated transferrin but with, however, similar or higher affinities for the helical domain than the C-lobe-competes with iron-saturated transferrin in an unknown state towards interaction with R1.  相似文献   

5.
6.
Asgeirsson B  Cekan P 《FEBS letters》2006,580(19):4639-4644
Temperature imposes limits on where life can thrive and this is evident in the evolution of the basic structural properties of proteins. Cold-adaptation of enzymes is one example, where the catalytic rate constant (k(cat)) is increased compared with hot-acclimated homologous under identical assay conditions. Trypsin I from Atlantic cod (Gadus morhua) has catalytic efficiency (k(cat)/K(m)) for amide hydrolysis that is 17-fold larger than observed for bovine trypsin. Here, the individual rate-constants for association of substrate (k(1)), dissociation of substrate (k(-1)), and acylation of the enzyme (k(2)) have been determined using benzoyl-Arg-p-nitroanilide or benzyloxycarbonyl-Gly-Pro-Arg-p-nitroanilide as substrates. Rather unexpectedly, by far the largest difference (37-fold increase) was observed in k(1), the rate constant for binding of substrate. The cold-adaptation of the dissociation and catalytic steps were not as prominent (increased by 3.7-fold). The length of substrate did have an effect by increasing the reaction rate by 70-fold, and again, the step most affected was the initial binding-step.  相似文献   

7.
The reaction of nitric oxide (NO) with fast and reduced cytochrome bo(3)(cyt bo(3)) from Escherichia coli has been investigated. The stoichiometry of NO binding to cyt bo(3) was determined using an NO electrode in the [NO] range 1-14 microM. Under reducing conditions, the initial decrease in [NO] following the addition of cyt bo(3) corresponded to binding of 1 NO molecule per cyt bo(3) functional unit. After this "rapid" NO binding phase, there was a slow, but significant rate of NO consumption ( approximately 0.3molNOmol bo(3)(-1)min(-1)), indicating that cyt bo(3) possesses a low level of NO reductase activity. The binding of NO to fast pulsed enzyme was also investigated. The results show that in the [NO] range used (1-14 microM) both fast and pulsed oxidised cyt bo(3) bind NO with a stoichiometry of 1:1 with an observed dissociation constant of K(d)=5.6+/-0.6 microM and that NO binding was inhibited by the presence of Cl(-). The binding of nitrite to the binuclear centre causes spectral changes similar to those observed upon NO binding to fast cyt bo(3). These results are discussed in relation to the model proposed by Wilson and co-workers [FEBS Lett. 414 (1997) 281] where the binding of NO to Cu(B)(II) results in the formation of the nitrosonium (Cu(B)(I)-NO(+)) complex. NO(+) then reacts with OH(-), a Cu(B) ligand, to form nitrite, which can bind at the binuclear centre. This work suggests for the first time that the binding of NO to oxidised cyt bo(3) does result in the reduction of Cu(B).  相似文献   

8.
9.
Fago A  Mathews AJ  Moens L  Dewilde S  Brittain T 《FEBS letters》2006,580(20):4884-4888
Previously identified, potentially neuroprotective reactions of neuroglobin require the existence of yet unknown redox partners. We show here that the reduction of ferric neuroglobin by cytochrome b(5) is relatively slow (k=6 x 10(2)M(-1)s(-1) at pH 7.0) and thus is unlikely to be of physiological significance. In contrast, the reaction between ferrous neuroglobin and ferric cytochrome c is very rapid (k=2 x 10(7)M(-1)s(-1)) with an apparent overall equilibrium constant of 1 microM. Based on this data we propose that ferrous neuroglobin may well play a role in preventing apoptosis.  相似文献   

10.
The general unireactant modifier mechanism in the absence of product can be described by the following linked reactions: E + S k1 in equilibrium k-1 ES k3----E + P; E + I k5 in equilibrium k-5 EI; EI + S k2 in equilibrium k-2 ESI k4----EI + P; and ES + I k6 in equilibrium k-6 ESI where S is a substrate and I is an effector. A full steady state treatment yields a velocity equation that is second degree in both [S] and [I]. Two different conditions (or assumptions) permit reduction of the velocity equation to one that is first degree in [S] and [I]. These are (a) that k-2k3 = k-1k4 (Frieden, C., J. Biol. Chem. 239, pp. 3522-3531, (1964)) and (b) that the I-binding reactions are at equilibrium (Reinhart, G. D., Arch. Biochem. Biophys. 224, pp. 389-401 (1983)). It is shown that each condition gives rise to the other (i.e., if the I-binding reactions are at equilibrium, then k-2k3 must equal k-1k4 and vice-versa). If one assumes equilibrium for the I-binding steps, the velocity equation derived by the method of Cha (J. Biol. Chem. 243, pp. 820-825 (1968)) is apparently second degree in [I] (Segel, I. H., Enzyme Kinetics, p. 838, Wiley-Interscience (1975)), but reduces to a first degree equation when the relationship derived by Frieden is inserted. If one starts by assuming a single equilibrium condition for I binding, e.g., k-5[EI] = k5[E][I] or k-6[ESI] = k6[ES][I], then a traditional algebraic manipulation of the remaining steady state equations provides first degree expressions for the concentrations of all enzyme species and also discloses the Frieden relationship.  相似文献   

11.
12.
The whole-cell voltage-clamp technique was used in rat cardiac myocytes to investigate the kinetics of ADP binding to phosphorylated states of Na,K-ATPase and its effects on presteady-state Na(+)-dependent charge movements by this enzyme. Ouabain-sensitive transient currents generated by Na,K-ATPase functioning in electroneutral Na(+)-Na(+) exchange mode were measured at 23 degrees C with pipette ADP concentrations ([ADP]) of up to 4.3 mM and extracellular Na(+) concentrations ([Na](o)) between 36 and 145 mM at membrane potentials (V(M)) from -160 to +80 mV. Analysis of charge-V(M) curves showed that the midpoint potential of charge distribution was shifted toward more positive V(M) both by increasing [ADP] at constant Na(+)(o) and by increasing [Na](o) at constant ADP. The total quantity of mobile charge, on the other hand, was found to be independent of changes in [ADP] or [Na](o). The presence of ADP increased the apparent rate constant for current relaxation at hyperpolarizing V(M) but decreased it at depolarizing V(M) as compared to control (no added ADP), an indication that ADP binding facilitates backward reaction steps during Na(+)-Na(+) exchange while slowing forward reactions. Data analysis using a pseudo three-state model yielded an apparent K(d) of approximately 6 mM for ADP binding to and release from the Na,K-ATPase phosphoenzyme; a value of 130 s(-1) for k(2), a rate constant that groups Na(+) deocclusion/release and the enzyme conformational transition E(1) approximately P --> E(2)-P; a value of 162 s(-1)M(-1) for k(-2), a lumped second-order V(M)-independent rate constant describing the reverse reactions; and a Hill coefficient of approximately 1 for Na(+)(o) binding to E(2)-P. The results are consistent with electroneutral release of ADP before Na(+) is deoccluded and released through an ion well. The same approach can be used to study additional charge-moving reactions and associated electrically silent steps of the Na,K-pump and other transporters.  相似文献   

13.
In Desulfovibrio metabolism, periplasmic hydrogen oxidation is coupled to cytoplasmic sulfate reduction via transmembrane electron transfer complexes. Type II tetraheme cytochrome c3 (TpII-c3), nine-heme cytochrome c (9HcA) and 16-heme cytochrome c (HmcA) are periplasmic proteins associated to these membrane-bound redox complexes and exhibit analogous physiological function. Type I tetraheme cytochrome c3 (TpI-c3) is thought to act as a mediator for electron transfer from hydrogenase to these multihemic cytochromes. In the present work we have investigated Desulfovibrio africanus (Da) and Desulfovibrio vulgaris Hildenborough (DvH) TpI-c3/TpII-c3 complexes. Comparative kinetic experiments of Da TpI-c3 and TpII-c3 using electrochemistry confirm that TpI-c3 is much more efficient than TpII-c3 as an electron acceptor from hydrogenase (second order rate constant k = 9 x 10(8) M(-1) s(-1), K(m) = 0.5 microM as compared to k = 1.7 x 10(7) M(-1) s(-1), K(m) = 40 microM, for TpI-c3 and TpII-c3, respectively). The Da TpI-c3/TpII-c3 complex was characterized at low ionic strength by gel filtration, analytical ultracentrifugation and cross-linking experiments. The thermodynamic parameters were determined by isothermal calorimetry titrations. The formation of the complex is mainly driven by a positive entropy change (deltaS = 137(+/-7) J mol(-1) K(-1) and deltaH = 5.1(+/-1.3) kJ mol(-1)) and the value for the association constant is found to be (2.2(+/-0.5)) x 10(6) M(-1) at pH 5.5. Our thermodynamic results reveal that the net increase in enthalpy and entropy is dominantly produced by proton release in combination with water molecule exclusion. Electrostatic forces play an important role in stabilizing the complex between the two proteins, since no complex formation is detected at high ionic strength. The crystal structure of Da TpI-c3 has been solved at 1.5 angstroms resolution and structural models of the complex have been obtained by NMR and docking experiments. Similar experiments have been carried out on the DvH TpI-c3/TpII-c3 complex. In both complexes, heme IV of TpI-c3 faces heme I of TpII-c3 involving basic residues of TpI-c3 and acidic residues of TpII-c3. A secondary interacting site has been observed in the two complexes, involving heme II of Da TpII-c3 and heme III of DvH TpI-c3 giving rise to a TpI-c3/TpII-c3 molar ratio of 2:1 and 1:2 for Da and DvH complexes, respectively. The physiological significance of these alternative sites in multiheme cytochromes c is discussed.  相似文献   

14.
Reaction between the dinuclear model hydrolases [M2(μ-OAc)2(OAc)2(μ-H2O)(tmen)2]; M = Ni (1); M = Co (2) and trimethylsilyltrifluoromethanesulphonate (TMS-OTf) under identical reaction conditions gives the mononuclear complex [Ni(OAc)(H2O)2(tmen)][OTf] · H2O (3) in the case of nickel and the dinuclear complex [Co2(μ-OAc)2(μ-H2O)2(tmen)2][OTf]2 (4) in the case of cobalt.Reaction of (3) with urea gives the previously reported [Ni(OAc)(urea)2(tmen)][OTf] (5), whereas (4) gives [Co2(OAc)3(urea)(tmen)2][OTf] (6) previously obtained by direct reaction of (2) with urea. Both (3) and (4) react with monohydroxamic acids (RHA) to give the dihydroxamate bridged dinuclear complexes [M2(μ-OAc)(μ-RA)2(tmen)2][OTf]; M = Ni (7); M = Co (8) previously obtained by the reaction of (1) and (2) with RHA, illustrating the greater ability of hydroxamic acids to stabilize dinuclear complexes over that of urea by means of their bridging mode, and offering a possible explanation for the inhibiting effect of hydroxamic acids by means of their displacing bridging urea in a possible intermediate invoked in the action of urease.  相似文献   

15.
In this work the interaction of Hydroxyzine, Promethazine and Thioridazine with Langmuir films of dipalmitoylphosphatidylcholine (dpPC) and dipalmitoylphosphatidic acid (dpPA), is studied. Temporal variations in lateral surface pressure (pi) were measured at different initial pi (pi(i)), subphase pH and drug-concentration. Drugs with the smallest (PRO) and largest (HYD) molecular size exhibited the lowest adsorption (k(a)) and the highest desorption (k(d)) rate constant values, respectively. The affinity binding constants (K(b)) obtained in monolayers followed the same profile (K(b,PRO) < K(b,HYD) < K(b,THI)) of the egg-PC/water partition coefficients (P) determined in bilayers. The drug concentration required to reach the half-maximal Deltapi at pi(i) = 14 mN/m (K(0.5)), was very sensitive to pH. The maximal increment in pi upon drug incorporation into the monolayer (deltapi(max)) will depend on the phospholipid collapse pressure (pi(c)), the monolayers's compressibility and drug's size, shape and charge. The higher pi(c) of dpPC lead to higher pi(cut-off) values (maximal pi allowing drug penetration), if compared with dpPA. In dpPC and dpPA pi(cut-off) decreased as a function of the molecular size of the uncharged drugs. In dpPA, protonated drugs became electrostatically trapped at the monolayer surface hence drug penetration, monolayer deformation and pi increase were impaired and the correlation between pi(cut-off) and drug molecular size was lost.  相似文献   

16.
Rhodobacter sphaeroides contains two enoyl-acyl carrier protein (ACP) reductases, FabI(1) and FabI(2). However, FabI(1) displays most of the cellular enzyme activity. The spontaneous diazaborine-resistant mutation was mapped as substitution of glutamine for proline 155 (P155Q) of FabI(1). The mutation of FabI(1)[P155Q] increased the specificity constants (k(cat)/K(m)) for crotonyl-ACP and NADH by more than 2-fold, while the site-directed mutation G95S (FabI(1)[G95S]), corresponding to the well-known G93 mutation of Escherichia coli FabI, rather decreased the values. Inhibition kinetics of the enzymes revealed that triclosan binds to the enzyme in the presence of NAD(+), while the diazaborine appears to interact with NADH and NAD(+) in the enzyme active site. The apparent inhibition constant K(i)(') of triclosan for FabI(1)[P155Q] and FabI(1)[G95S] at saturating NAD(+) were approximately 80- and 3-fold higher than that for the wild-type enzyme, respectively, implying that the inhibition was remarkably impaired by the P155Q mutation. The similar levels of K(i)(') of diazaborine for the mutant enzymes were also observed with respect to NAD(+). Thus, the novel mutation P155Q appears to disturb the binding of inhibitors to the enzyme without affecting the catalytic efficiency.  相似文献   

17.
Y-family DNA polymerases (DNAPs) are often required in cells to synthesize past DNA-containing lesions, such as [+ ta]-B[a]P-N2-dG, which is the major adduct of the potent mutagen/carcinogen benzo[a]pyrene. The current model for the non-mutagenic pathway in Escherichia coli involves DNAP IV inserting deoxycytidine triphosphate opposite [+ ta]-B[a]P-N2-dG and DNAP V doing the next step(s), extension. We are investigating what structural differences in these related Y-family DNAPs dictate their functional differences. X-ray structures of Y-family DNAPs reveal a number of interesting features in the vicinity of the active site, including (1) the “roof-amino acid” (roof-aa), which is the amino acid that lies above the nucleobase of the deoxynucleotide triphosphate (dNTP) and is expected to play a role in dNTP insertion efficiency, and (2) a cluster of three amino acids, including the roof-aa, which anchors the base of a loop, whose detailed structure dictates several important mechanistic functions. Since no X-ray structures existed for UmuC (the polymerase subunit of DNAP V) or DNAP IV, we previously built molecular models. Herein, we test the accuracy of our UmuC(V) model by investigating how amino acid replacement mutants affect lesion bypass efficiency. A ssM13 vector containing a single [+ ta]-B[a]P-N2-dG is transformed into E. coli carrying mutations at I38, which is the roof-aa in our UmuC(V) model, and output progeny vector yield is monitored as a measure of the relative efficiency of the non-mutagenic pathway. Findings show that (1) the roof-aa is almost certainly I38, whose β-carbon branching R-group is key for optimal activity, and (2) I38/A39/V29 form a hydrophobic cluster that anchors an important mechanistic loop, aa29-39. In addition, bypass efficiency is significantly lower both for the I38A mutation of the roof-aa and for the adjacent A39T mutation; however, the I38A/A39T double mutant is almost as active as wild-type UmuC(V), which probably reflects the following. Y-family DNAPs fall into several classes with respect to the [roof-aa/next amino acid]: one class has [isoleucine/alanine] and includes UmuC(V) and DNAP η (from many species), while the second class has [alanine (or serine)/threonine] and includes DNAP IV, DNAP κ (from many species), and Dpo4. Thus, the high activity of the I38A/A39T double mutant probably arises because UmuC(V) was converted from the V/η class to the IV/κ class with respect to the [roof-aa/next amino acid]. Structural and mechanistic aspects of these two classes of Y-family DNAPs are discussed.  相似文献   

18.
Considering the dynamic nature of CYPs, methods that reveal information about substrate and enzyme dynamics are necessary to generate predictive models. To compare substrate dynamics in CYP2E1 and CYP2A6, intramolecular isotope effect experiments were conducted, using deuterium labeled substrates: o-xylene, m-xylene, p-xylene, 2,6-dimethylnaphthalene, and 4,4'-dimethylbiphenyl. Competitive intermolecular experiments were also conducted using d(0)- and d(6)-labeled p-xylene. Both CYP2E1 and CYP2A6 displayed full isotope effect expression for o-xylene oxidation and almost complete suppression for dimethylbiphenyl. Interestingly, (k(H)/k(D))(obs) for d(3)-p-xylene oxidation ((k(H)/k(D))(obs)=6.04 and (k(H)/k(D))(obs)=5.53 for CYP2E1 and CYP2A6, respectively) was only slightly higher than (k(H)/k(D))(obs) for d(3)-dimethylnaphthalene ((k(H)/k(D))(obs)=5.50 and (k(H)/k(D))(obs)=4.96, respectively). One explanation is that in some instances (k(H)/k(D))(obs) values are generated by the presence of two substrates-bound simultaneously to the CYP. Speculatively, if this explanation is valid, then intramolecular isotope effect experiments should be useful in the mechanistic investigation of P450 cooperativity.  相似文献   

19.
Among the mechanisms proposed for the increase in discharge of sino-atrial node (SAN) by norepinephrine (NE) are an increase in the hyperpolarization-activated current I(f) and in the slow inward current I(Ca,L). If I(f) is the primary mechanism, cesium (a blocker of I(f)) should eliminate the positive chronotropic effect of NE. If I(Ca,L), is involved, [Ca(2+)](o) should condition NE effects. We studied the electrophysiological changes induced by NE in isolated guinea pig SAN superfused in vitro with Tyrode solution (both SAN dominant and subsidiary pacemaker mechanisms are present) as well as with high [K(+)](o), higher Cs(+) or Ba(2+) (only the dominant pacemaker mechanism is present). In Tyrode solution, NE (0.5-1microM) increased the SAN rate and adding Cs(+) (approximately 12 mM) caused a decaying voltage tail during diastole in subsidiary pacemakers. NE enhanced the Cs(+)-induced tail, and increased the rate but less than in Tyrode solution. In higher [Cs(+)](o) (15- 18 mM), Ba(2+) (1 mM) or Ba(2+) plus Cs(+) (10 mM) dominant action potentials (not followed by a tail) were present and NE accelerated them as in Tyrode solution. In high [K(+)](o), NE increased the rate in the absence and presence of Cs(+), Ba(2+) or Ba(2+) plus Cs(+). In these solutions, NE increased the overshoot and maximum diastolic potential of dominant action potentials (APs) and increased the rate by steepening diastolic depolarization and shifting the threshold for upstroke to more negative values. High [Ca(2+)](o) alone increased the rate and NE enhanced this action, whereas low [Ca(2+)](o) reduced or abolished the increase in rate by NE. In SAN quiescent in high [K(+)](o) plus indapamide, NE induced spontaneous discharge by decreasing the resting potential and initiating progressively larger voltage oscillations. Thus, NE increases the SAN rate by acting primarily on dominant APs in a manner consistent with an increase of I(Ca,L) and I(K) and under conditions where I(f) is either blocked or not activated. NE INITIATES spontaneous discharge by inducing voltage oscillations unrelated to I(f).  相似文献   

20.
Popovic S  Urbán E  Lukic M  Conlon JM 《Peptides》2012,34(2):275-282
The pathogenesis of acne vulgaris is multifactorial involving infection of the pilosebaceous unit with Propionibacterium acnes and a cytokine-mediated inflammatory response. Five frog skin-derived antimicrobial peptides ([D4k]ascaphin-8, [G4K]XT-7, [T5k]temporin-DRa, brevinin-2GU, and B2RP-ERa), chosen for their low hemolytic activity against human erythrocytes, were assessed for their effects on the growth of clinical isolates of P. acnes and on the release of pro-inflammatory and anti-inflammatory cytokines from peripheral blood mononuclear (PBM) cells. All peptides inhibited the growth of P. acnes with the highest potency exhibited by [D4k]ascaphin-8 (minimum inhibitory concentration, MIC=3-12.5 μM). Release of TNF-α from concanavalin A (ConA)-stimulated PBM cells was significantly reduced by [D4k]ascaphin-8, [G4K]XT-7, brevinin-2GU, and B2RP-ERa (1 and 20 μg/ml) and by [T5k]temporin-DRa (20 μg/ml). Release of IFN-γ from unstimulated PBM cells was significantly reduced by [D4k]ascaphin-8 and brevinin-2GU (1 and 20 μg/ml). No peptide showed significant effects on Il-17 release. Release of the anti-inflammatory cytokines TGF-β, IL-4, and IL-10 from both unstimulated and ConA-treated PBM cells was significantly increased by [T5k]temporin-DRa and B2RP-ERa (1 and 20μg/ml). The potent activities of [D4k]ascaphin-8 and [T5k]temporin-DRa in inhibiting the growth of P. acnes and the release of pro-inflammatory cytokines, and in stimulating the release of anti-inflammatory cytokines suggest a possible therapeutic role in the treatment of acne vulgaris.  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