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 250 毫秒
1.
Understanding the mechanisms that dictate protein stability is of large relevance, for instance, to enable design of temperature-tolerant enzymes with high enzymatic activity over a broad temperature interval. In an effort to identify such mechanisms, we have performed a detailed comparative study of the folding thermodynamics and kinetics of the ribosomal protein S16 isolated from a mesophilic (S16meso) and hyperthermophilic (S16thermo) bacterium by using a variety of biophysical methods. As basis for the study, the 2.0 Å X-ray structure of S16thermo was solved using single wavelength anomalous dispersion phasing. Thermal unfolding experiments yielded midpoints of 59 and 111 °C with associated changes in heat capacity upon unfolding (ΔCp0) of 6.4 and 3.3 kJ mol− 1 K− 1, respectively. A strong linear correlation between ΔCp0 and melting temperature (Tm) was observed for the wild-type proteins and mutated variants, suggesting that these variables are intimately connected. Stopped-flow fluorescence spectroscopy shows that S16meso folds through an apparent two-state model, whereas S16thermo folds through a more complex mechanism with a marked curvature in the refolding limb indicating the presence of a folding intermediate. Time-resolved energy transfer between Trp and N-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-yl)methyl iodoacetamide of proteins mutated at selected positions shows that the denatured state ensemble of S16thermo is more compact relative to S16meso. Taken together, our results suggest the presence of residual structure in the denatured state ensemble of S16thermo that appears to account for the large difference in quantified ΔCp0 values and, in turn, parts of the observed extreme thermal stability of S16thermo. These observations may be of general importance in the design of robust enzymes that are highly active over a wide temperature span.  相似文献   

2.
Polcalcins are small EF‐hand proteins believed to assist in regulating pollen‐tube growth. Phl p 7, from timothy grass (Phleum pratense), crystallizes as a domain‐swapped dimer at low pH. This study describes the solution structures of the recombinant protein in buffered saline at pH 6.0, containing either 5.0 mM EDTA, 5.0 mM Mg2+, or 100 μM Ca2+. Phl p 7 is monomeric in all three ligation states. In the apo‐form, both EF‐hand motifs reside in the closed conformation, with roughly antiparallel N‐ and C‐terminal helical segments. In 5.0 mM Mg2+, the divalent ion is bound by EF‐hand 2, perturbing interhelical angles and imposing more regular helical structure. The structure of Ca2+‐bound Phl p 7 resembles that previously reported for Bet v 4—likewise exposing apolar surface to the solvent. Occluded in the apo‐ and Mg2+‐bound forms, this surface presumably provides the docking site for Phl p 7 targets. Unlike Bet v 4, EF‐hand 2 in Phl p 7 includes five potential anionic ligands, due to replacement of the consensus serine residue at –x (residue 55 in Phl p 7) with aspartate. In the Phl p 7 crystal structure, D55 functions as a helix cap for helix D. In solution, however, D55 apparently serves as a ligand to the bound Ca2+. When Mg2+ resides in site 2, the D55 carboxylate withdraws to a distance consistent with a role as an outer‐sphere ligand. 15N relaxation data, collected at 600 MHz, indicate that backbone mobility is limited in all three ligation states. Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

3.
Many macromolecular interactions, including protein‐nucleic acid interactions, are accompanied by a substantial negative heat capacity change, the molecular origins of which have generated substantial interest. We have shown previously that temperature‐dependent unstacking of the bases within oligo(dA) upon binding to the Escherichia coli SSB tetramer dominates the binding enthalpy, ΔHobs, and accounts for as much as a half of the observed heat capacity change, ΔCp. However, there is still a substantial ΔCp associated with SSB binding to ssDNA, such as oligo(dT), that does not undergo substantial base stacking. In an attempt to determine the origins of this heat capacity change, we have examined by isothermal titration calorimetry (ITC) the equilibrium binding of dT(pT)34 to SSB over a broad pH range (pH 5.0–10.0) at 0.02 M, 0.2 M NaCl and 1 M NaCl (25°C), and as a function of temperature at pH 8.1. A net protonation of the SSB protein occurs upon dT(pT)34 binding over this entire pH range, with contributions from at least three sets of protonation sites (pKa1 = 5.9–6.6, pKa2 = 8.2–8.4, and pKa3 = 10.2–10.3) and these protonation equilibria contribute substantially to the observed ΔH and ΔCp for the SSB‐dT(pT)34 interaction. The contribution of this coupled protonation (∼ −260 to −320 cal mol−1 K−1) accounts for as much as half of the total ΔCp. The values of the “intrinsic” ΔCp,0 range from −210 ± 33 cal mol−1 °K−1 to −237 ± 36 cal mol−1K−1, independent of [NaCl]. These results indicate that the coupling of a temperature‐dependent protonation equilibria to a macromolecular interaction can result in a large negative ΔCp, and this finding needs to be considered in interpretations of the molecular origins of heat capacity changes associated with ligand‐macromolecular interactions, as well as protein folding. Proteins 2000;Suppl 4:8–22. © 2000 Wiley‐Liss, Inc.  相似文献   

4.
Several experimental methods (circular dichroism, viscosity, intrinsic fluorescence, and fluorescence labeling) were used to study the conformational folding/unfolding transitions in a compact monomeric form of the Caf113-149 subunit under the action of guanidine hydrochloride in the temperature range 5–45°C. It has been shown that transitions always occur between two major states (unfolded and compact). This has made it possible to determine all the main thermodynamic functions that characterize the compact state of the Caf113-149 subunit: stability temperature T m, free energy of stabilization ΔG st, enthalpy ΔH tr, and heat capacity jump ΔC in collapse of the structure. These data have been confirmed by an independent experiment on melting of fluorescently labeled protein.  相似文献   

5.
Despite its importance as a target in anti-cancer therapeutics and the numerous rational-based inhibitor design efforts aimed at it, there are only limited data available on structural-thermodynamic relationships of interactions of the N-terminal ATP-binding domain of Hsp90 (N-Hsp90). Here, we redress this by presenting an investigation of binding of nucleotides and ansamycin compounds to this domain. Interactions of nucleotides with N-Hsp90 are relatively weak (> 10 μM) and are strongly enthalpy driven over the temperature range 10-25 °C. Geldanamycin (GA) and its analogues 17-AAG [17-(allylamino)-17-demethoxy-GA] and 17-DMAG (17-N,N-dimethylaminoethylamino-17-demethoxy-GA) bind more strongly and have a dominant favourable enthalpic contribution over the temperature range investigated. We investigated the temperature dependence of the enthalpic contribution to binding. We found that while the ansamycin compounds have the commonly observed negative value, the nucleotides show a negligible or even a positive ΔCp of binding. These data represent the first observation of a single binding site for which interactions with different ligands result in both negative and positive ΔCp values. By addressing the likely impact of the potential contributions from protein-ligand interactions, we are able to attribute the anomalous ΔCp for the nucleotides largely to a change in the conformation of the domain structure and local motion in the lid region of N-Hsp90 with the concomitant exposure of hydrophobic amino acid side chains.  相似文献   

6.
The interactions between the HIV Rev-responsive element (RRE) RNA and the HIV regulatory protein Rev, are crucial for the HIV life-cycle. Earlier, we showed that single C2H2 zinc fingers (znfs) have the same binding site as the Rev peptide and exhibit nanomolar affinity. In this study, the specific role of amino acid side chains and molecular processes involved with complex formation were investigated by perturbation of the binding energetics via changes in temperature, pH, buffers, and salt concentrations, as well as znf and RNA mutations, by isothermal titration calorimetry. Interestingly, despite the large cationic charge on the znfs, the number of interactions with the RNA phosphate backbone was lower than intuitively expected. The presence of binding induced protonation was established by ITC and localized by NMR to a histidine on the znf β-sheet. The ΔCp of znf-RNA binding was observed to be substantially negative and could not be accounted for by conventional solvent-accessible surface area models. An alternative model, based on the extent of hydrogen bond changes as a result of differences in ligand-induced water displacement at the binding site, provided reasonable explanation of the trends in ΔCp, as well as ΔH and ΔS. Our studies show that incorporation of favorable interactions at the solvent-excluded binding interface can be used to alleviate the unfavorable enthalpic penalties of displacing water molecules from the hydrated RNA surface.  相似文献   

7.
Electrostatic interactions have a central role in some biological processes, such as recognition of charged ligands by proteins. We characterized the binding energetics of yeast triosephosphate isomerase (TIM) with phosphorylated inhibitors 2-phosphoglycollate (2PG) and phosphoglycolohydroxamate (PGH). We determined the thermodynamic parameters of the binding process (Kb, ΔGb, ΔHb, ΔSb and ΔCp) with different concentrations of NaCl, using fluorimetric and calorimetric titrations in the conventional mode of ITC and a novel method, multithermal titration calorimetry (MTC), which enabled us to measure ΔCp in a single experiment. We ruled out specific interactions of Na+ and Cl- with the native enzyme and did not detect significant linked protonation effects upon the binding of inhibitors. Increasing ionic strength (I) caused Kb, ΔGb and ΔHb to become less favorable, while ΔSb became less unfavorable. From the variation of Kb with I, we determined the electrostatic contribution of TIM−2PG and TIM−PGH to ΔGb at I = 0.06 M and 25 °C to be 36% and 26%, respectively. The greater affinity of PGH for TIM is due to a more favorable ΔHb compared to 2PG (by 19-24 kJ mol-1 at 25 °C). This difference is compatible with PGH establishing up to five more hydrogen bonds with TIM. Both binding ΔCps were negative, and less negative with increasing ionic strength. ΔCps at I = 0.06 M were much more negative than predicted by surface area models. Water molecules trapped in the interface when ligands bind to protein could explain the highly negative ΔCps. Thermodynamic binding functions for TIM−2PG changed more with ionic strength than those for TIM−PGH. This greater dependence is consistent with linked, but compensated, protonation equilibriums yielding the dianionic species of 2PG that binds to TIM, process that is not required for PGH.  相似文献   

8.
Fluorescence and circular dichroism data as a function of temperature were obtained to characterize the unfolding of nuclease A and two of its less stable mutants. These spectroscopic data were obtained with a modified instrument that enables the nearly simultaneous detection of both fluorescence and CD data on the same sample. A global analysis of these multiple datasets yielded an excellent fit of a model that includes a change in the heat capacity change, ΔCp, between the unfolded and native states. This analysis gives a ΔCp of 2.2 kcal/mol/·K for thermal unfolding of the WT protein and 1.3 and 1.8 kcal/mol/K for the two mutants. These ΔCp values are consistent with significant population of the cold unfolded state at ∼0°C. Independent evidence for the existence of a cold unfolded state is the observation of a separately migrating peak in size exclusion chromatography. The new chromatographic peak is seen near 0°C, has a partition coefficient corresponding to a larger hydrodynamic radius, and shows a red-shifted fluorescence spectrum, as compared to the native protein. Data also indicate that the high-temperature unfolded form of mutant nuclease is relatively compact. Size exclusion chromatography shows the high temperature unfolded form to have a hydrodynamic radius that is larger than that for the native form, but smaller than that for the urea or pH-induced unfolded forms. Addition of chemical denaturants to the high-temperature unfolded form causes a further unfolding of the protein, as indicated by an increase in the apparent hydrodynamic radius and a decrease in the rotational correlation time for Trp140 (as determined by fluorescence anisotropy decay measurements). Proteins 28:227–240, 1997 © 1997 Wiley-Liss Inc.  相似文献   

9.
《Inorganica chimica acta》1988,145(2):267-271
Detailed synthetic and mechanistic studies of the addition of 2, 6-dimethylaniline to the organometallic complexes [Fe(CO)3(1–5-η-dienyl)]BF4 (1, dienyl=C6H7 or C7H9) indicates the general rate law kobs=ka [2,6-(Me)2C6H3NH2]+kb which is consistent with an equilibrium process. The greater reactivity of the C6H7 complex and the low ΔHa3 and large negative ΔSa3 values are in accordance with direct addition of the amine to the dienyl rings of 1. On the other hand the relatively much higher ΔHb 3 values are consistent with bond cleavage in dissociation as is the positive ΔSb 3 value of +220 J K−1 mol−1 determined for the C6H7 reaction. The negative ΔSb3 value of −43 J K−1 mol−1 found for the C7H9 reaction suggests the presence of an ordered transition state through which the starting dienyl complex is reformed via some internal SN2 process.  相似文献   

10.
Eukaryotic and archaeal translation initiation factors 2, heterotrimers that consist of α-, β-, and γ-subunits, deliver methionylated initiator tRNA to a small ribosomal subunit in a manner that depends on GTP. To evaluate correlation of the function and association of the subunits, we used isothermal titration calorimetry to analyze the thermodynamics of the interactions between the α- and γ-subunits in the presence or absence of a nonhydrolyzable GTP analog or GDP. The α-subunits bound to the γ-subunit with large heat capacity change (ΔCp) values. The ΔH and ΔCp values for the interaction between the α- and γ-subunits varied in the presence of the GTP analog but not in the presence of GDP. These results suggest that the binding of both the α-subunit and GTP changes the conformation of the switch region of the γ-subunit and increases the affinity of the γ-subunit for tRNA.  相似文献   

11.
Calorimetric and fluorescence techniques were used to characterize the binding of aminoglycosides-neomycin, paromomycin, and ribostamycin, with 5′-dA12-x-dT12-x-dT12-3′ intramolecular DNA triplex (x = hexaethylene glycol) and poly(dA)·2poly(dT) triplex. Our results demonstrate the following features: (1) UV thermal analysis reveals that the Tm for triplex decreases with increasing pH value in the presence of neomycin, while the Tm for the duplex remains unchanged. (2) The binding affinity of neomycin decreases with increased pH, although there is an increase in observed binding enthalpy. (3) ITC studies conducted in two buffers (sodium cacodylate and MOPS) yield the number of protonated drug amino groups (Δn) as 0.29 and 0.40 for neomycin and paromomycin interaction with 5′-dA12-x-dT12-x-dT12-3′, respectively. (4) The specific heat capacity change (ΔCp) determined by ITC studies is negative, with more negative values at lower salt concentrations. From 100 mM to 250 mM KCl, the ΔCp ranges from −402 to −60 cal/(mol K) for neomycin. At pH 5.5, a more positive ΔCp is observed, with a value of −98 cal/(mol K) at 100 mM KCl. ΔCp is not significantly affected by ionic strength. (5) Salt dependence studies reveal that there are at least three amino groups of neomycin participating in the electrostatic interactions with the triplex. (6) FID studies using thiazole orange were used to derive the AC50 (aminoglycoside concentration needed to displace 50% of the dye from the triplex) values. Neomycin shows a seven fold higher affinity than paromomycin and eleven fold higher affinity than ribostamycin at pH 6.8. (7) Modeling studies, consistent with UV and ITC results, show the importance of an additional positive charge in triplex recognition by neomycin. The modeling and thermodynamic studies indicate that neomycin binding to the DNA triplex depends upon significant contributions from charge as well as shape complementarity of the drug to the DNA triplex Watson–Hoogsteen groove.  相似文献   

12.
《Inorganica chimica acta》1988,142(2):291-299
In coordinating solvents, the complex 1, 4, 8, 11- tetramethyl-1, 4, 8, 11-tetraazacyclotetradecane nickel(II) bisperchlorate exists as an equilibrium mixture involving four coordinate R,S,R,S-[Ni(tmc)]2+ and five coordinate R,S,R,S-[Ni(tmc)(solvent)]2+ species. Spectrophotometric measurements of this equilibrium in a number of solvents have been conducted over a range of temperatures and pressures. The stability order for the five coordinate complex in the solvents investigated is CH3CN>DMF>DMSO>C6H5CN> H2O>ClCH2CN at 25 °C. Differences in stability are considered in terms of the measured thermodynamic parameters ΔH° and ΔS°. Both steric and electronic factors were found to influence solvent coordination with the macrocyclic complex.For the equilibrium in CH3CN, C6H5CN, DMF and H2O, reaction volumes, ΔV°, of −3.2±0.5, −4.2±0.5, −0.2±0.5 and −0.5±0.5 cm3 mol−1 respectively have been determined. Each is significantly smaller than the corresponding solvent molar volume. The ΔV° for the equilibrium in CH3CN is comparable with the previously determined activation volume for exchange of this solvent on R, S, R, S- [Ni(tmc)(CH3CN)]2+. The equilibrium and measured volume parameters are discussed in relation to the mechanism for solvent exchange.  相似文献   

13.
《Insect Biochemistry》1989,19(8):759-765
Newly-laid eggs of the house cricket Acheta domesticus contain significant amounts of apolar ecdysteroid conjugates, which can be hydrolysed by prolonged incubation with a mixture of Helix pomatia gut hydrolases. The ecdysteroid released on hydrolysis of the apolar conjugates has been purified and identified as ecdysone by co-chromatography on normal-phase and reversed-phase HPLC and by fast-atom bombardment mass spectrometry.Starting with only 22 g newly-laid eggs (containing 16 μg conjugated ecdysone), the ecdysone conjugates have been purified by open column chromatography and four successive HPLC purification steps to give essentially pure apolar conjugates with a yield of 57%. The conjugates are shown to be a mixture of ecdysone 22-fatty acyl esters by co-chromatography with authentic reference compounds and by fast-atom bombardment mass spectrometry. The fatty acyl composition of the conjugates is very similar to that produced by the ovaries of A. domesticus from [3H]ecdysone in vitro (Whiting and Dinan, Biochem. J.252, 95–103, 1988). The major fatty acyl esters are the 22-palmitate (C16:0), 22-oleate (C18:1) and 22-linoleate (C18:2), with smaller amounts of the myristate (C14:0), stearate (C18:0) and arachidate (C20:0) esters.This report constitutes the first identification from an insect source of endogenous ecdysteroid 22-fatty acyl esters, which have previously been identified in ticks and as metabolites of exogenous [3H]ecdysone in several arthropod species.  相似文献   

14.
Thioredoxins are small, ubiquitous redox enzymes that reduce protein disulfide bonds by using a pair of cysteine residues present in a strictly conserved WCGPC catalytic motif. The Escherichia coli cytoplasm contains two thioredoxins, Trx1 and Trx2. Trx2 is special because it is induced under oxidative stress conditions and it has an additional N-terminal zinc-binding domain. We have determined the redox potential of Trx2, the pKa of the active site nucleophilic cysteine, as well as the stability of the oxidized and reduced form of the protein. Trx2 is more oxidizing than Trx1 (-221 mV versus -284 mV, respectively), which is in good agreement with the decreased value of the pKa of the nucleophilic cysteine (5.1 versus 7.1, respectively). The difference in stability between the oxidized and reduced forms of an oxidoreductase is the driving force to reduce substrate proteins. This difference is smaller for Trx2 (ΔΔG°H2O = 9 kJ/mol and ΔTm = 7. 4 °C) than for Trx1 (ΔΔG°H2O = 15 kJ/mol and ΔTm = 13 °C). Altogether, our data indicate that Trx2 is a significantly less reducing enzyme than Trx1, which suggests that Trx2 has a distinctive function. We disrupted the zinc center by mutating the four Zn2+-binding cysteines to serine. This mutant has a more reducing redox potential (-254 mV) and the pKa of its nucleophilic cysteine shifts from 5.1 to 7.1. The removal of Zn2+ also decreases the overall stability of the reduced and oxidized forms by 3.2 kJ/mol and 5.8 kJ/mol, respectively. In conclusion, our data show that the Zn2+-center of Trx2 fine-tunes the properties of this unique thioredoxin.  相似文献   

15.
The ΔF508 mutation in nucleotide-binding domain 1 (NBD1) of the cystic fibrosis transmembrane conductance regulator (CFTR) is the predominant cause of cystic fibrosis. Previous biophysical studies on human F508 and ΔF508 domains showed only local structural changes restricted to residues 509-511 and only minor differences in folding rate and stability. These results were remarkable because ΔF508 was widely assumed to perturb domain folding based on the fact that it prevents trafficking of CFTR out of the endoplasmic reticulum. However, the previously reported crystal structures did not come from matched F508 and ΔF508 constructs, and the ΔF508 structure contained additional mutations that were required to obtain sufficient protein solubility. In this article, we present additional biophysical studies of NBD1 designed to address these ambiguities. Mass spectral measurements of backbone amide 1H/2H exchange rates in matched F508 and ΔF508 constructs reveal that ΔF508 increases backbone dynamics at residues 509-511 and the adjacent protein segments but not elsewhere in NBD1. These measurements also confirm a high level of flexibility in the protein segments exhibiting variable conformations in the crystal structures. We additionally present crystal structures of a broader set of human NBD1 constructs, including one harboring the native F508 residue and others harboring the ΔF508 mutation in the presence of fewer and different solubilizing mutations. The only consistent conformational difference is observed at residues 509-511. The side chain of residue V510 in this loop is mostly buried in all non-ΔF508 structures but completely solvent exposed in all ΔF508 structures. These results reinforce the importance of the perturbation ΔF508 causes in the surface topography of NBD1 in a region likely to mediate contact with the transmembrane domains of CFTR. However, they also suggest that increased exposure of the 509-511 loop and increased dynamics in its vicinity could promote aggregation in vitro and aberrant intermolecular interactions that impede trafficking in vivo.  相似文献   

16.
A nondetergent sulfobetaine (NDSB) was found to improve unfolding reversibility of several proteins by inhibiting heat-induced aggregation. As a consequence, ΔHcalHvH ratios were also improved to values close to 1 for a two-state unfolding. NDSB is effective in a wide range of pH values and especially at acidic pH generally used to calculate ΔCp values by the Kirchhoff relation. The sulfobetaine also allows recording protein refolding by protecting the heat-induced unfolded state against aggregation.  相似文献   

17.
Thermodynamic parameters for the unfolding of as well as for the binding of Ca2+ to goat α-lactalbumin (GLA) and bovine α-lactalbumin (BLA) are deduced from isothermal titration calorimetry in a buffer containing 10 mM Tris-HCl, pH 7.5 near 25°C. Among the different parameters available, the heat capacity increments (ΔCp) offer the most direct information for the associated conformational changes of the protein variants. The ΔCp values for the transition from the native to the molten globule state are rather similar for both proteins, indicating that the extent of the corresponding conformational change is nearly identical. However, the respective ΔCp values for the binding of Ca2+ are clearly different. The data suggest that a distinct protein region is more sensitive to a Ca2+-dependent conformational change in BLA than is the case in GLA. By analysis of the tertiary structure we observed an extensive accumulation of negatively charged amino acids near the Ca2+-binding site of BLA. In GLA, the cluster of negative charges is reduced by the substitution of Glu-11 by Lys. The observed difference in ΔCp values for the binding of Ca2+ is presumably in part related to this difference in charge distribution.  相似文献   

18.
Maurotoxin (MTX) is a 34‐residue toxin that was isolated initially from the venom of the scorpion Scorpio maurus palmatus. Unlike the other toxins of the α‐KTx6 family (Pi1, Pi4, Pi7, and HsTx1), MTX exhibits a unique disulfide bridge organization of the type C1? C5, C2? C6, C3? C4, and C7? C8 (instead of the conventional C1? C5, C2? C6, C3? C7, and C4? C8, herein referred to as Pi1‐like) that does not prevent its folding along the classic α/β scaffold of scorpion toxins. MTXPi1 is an MTX variant with a conventional pattern of disulfide bridging without any primary structure alteration of the toxin. Here, using MTX and/or MTXPi1 as models, we investigated how the type of folding influences toxin recognition of the Shaker B potassium channel. Amino acid residues of MTX that were studied for Shaker B recognition were selected on the basis of their homologous position in charybdotoxin, a three disulfide‐bridged scorpion toxin also active on this channel type. These residues favored either an MTX‐ or MTXPi1‐like folding. Our data indicate clearly that Lys23 and Tyr32 (two out of ten amino acid residues studied) are the most important residues for Shaker B channel blockage by MTX. For activity on SKCa channels, the same amino acid residues also affect, directly or indirectly, the recognition of SK channels. The molecular modeling technique and computed docking indicate the existence of a correlation between the half cystine pairings of the mutated analogs and their activity on the Shaker B K+ channel. Overall, mutations in MTX could, or could not, change the reorganization of disulfide bridges of this molecule without affecting its α/β scaffold. However, changing of the peptide backbone (cross linking disulfide bridges from MTX‐like type vs MTXPi1‐like type) appears to have less impact on the molecule activity than mutation of certain key amino acids such as Lys23 and Tyr32 in this toxin. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

19.
The major components (50%) of the surface lipid extract of fungal spores (5.6% of dry spore wt) of Sphaerotheca fuliginea are esters of primary alcohols and fatty acids. Esters (15%) of primary alcohols and a Δ2t acid are present. The major acid moieties of the alkyl esters are C22 and C24 and of the Δ2t alkyl ester is Δ2t C22; for both classes eicosanol is the major primary alcohol. The major ester of each class was concluded to be eicosanyl docosanoate and eicosanyl trans-2-docosenoate. Minor components are saturated and Δ2t methyl and diol diesters and free fatty acids. The major acid moieties of the diol diesters are C22 and C24 and the major diol is 1,12-dodecanediol.  相似文献   

20.
Previousin vivostudies involving sequence 5′-CCCG1G2G3-3′ (SmaI site) have demonstrated that adducts ofN-2-acetylaminofluorene (AAF) to any of the three guanine residues of theSmaI sequence induce, with different efficiencies, two classes of −1 frameshift events, namely −G and −C mutations, referred to as targeted and semitargeted mutations, respectively. It has been proposed that both events occur during replication as a consequence of slippage events involving slipped mutagenic intermediates (SMIs). In order to evaluate the potential role of the UvrABC excinuclease in frameshift mutagenesis, we have studied the interaction of this enzyme with DNA molecules mimicking SMIsin vitro.In all of our constructions, when present, the AAF adduct was located on the third guanine residue of theSmaI site (5′-CCCG1G2G3-3′). This strand was referred to as the top strand, the complementary strand being the bottom strand. Double-stranded heteroduplexes mimicking the targeted and semitargeted SMIs contained a deletion of a C and a G within theSmaI sequence in the bottom strand and were designated ΔC/3 and ΔG/3 when modified with the AAF on the third guanine residue in the top strand or ΔC/O and ΔG/O when unmodified. The modified homoduplex was designatedSmaI/3.ΔC/O and ΔG/O were weakly recognized by UvrA2B, but not incised. All three AAF-modified substrates were recognized with similar efficiency and much more efficiently than unmodified heteroduplexes. With AAF-monomodified substrates, dissociation of UvrA2from the UvrA2B- DNA complex occurred more readily in heteroduplexes than in the homoduplex.SmaI/3 and ΔC/3 were incised with equal efficiency, while ΔG/3 was less incised. The position of the AAF lesion dictated the position of the incised phosphodiester bonds, suggesting that the presence of a bulge can modulate the yield but not the incision pattern of AAF-modified substrates. The finding that UvrABC excinuclease acts on substrates that mimic SMIs suggests that the nucleotide excision repair pathway may help in fixing frameshift mutations before the following round of replication.  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