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.
The action of the relaxing agent dantrolene on dipalmitoylphosphatidylcholine (DPPC) model membranes in the presence and absence of the general anesthetic halothane has been investigated by DSC and 31P-NMR. Dantrolene has a weak effect on both the thermodynamic and NMR parameters of the pure model membrane. When halothane is present in the system, the relaxing agent acts to counterbalance the strong anesthetic-induced membrane perturbation. This is reflected in DSC experiments by a change of the enthalpy variation (delta H) and of the main gel-to-fluid phase transition temperature (Tc) towards the values of the pure lipid system. The amount of halothane-induced small tumbling vesicles, as detected by 31P-NMR by the superposition of an isotropic line on a lamellar-type powder spectrum, is considerably reduced upon dantrolene addition. This means that the relaxing agent "cures" the membrane de-structuring action promoted by halothane. Membranes first treated with dantrolene are also protected from the halothane perturbation. So, the relaxing agent is both "curative" and "preventative" against halothane. The optimum effect is obtained for 1 dantrolene molecule per ca 34 halothane molecules. The mechanisms of action were discussed in relation to membrane fluidity.  相似文献   

2.
3.
The 13-residue cathelicidins indolicidin and tritrpticin are part of a group of relatively short tryptophan-rich antimicrobial peptides that hold potential as future substitutes for antibiotics. Differential scanning calorimetry (DSC) has been applied here to study the effect of indolicidin and tritrpticin as well as five tritrpticin analogs on the phase transition behaviour of model membranes made up of zwitterionic dimyristoylphosphatidylcholine (DMPC, DMPC/cholesterol) and anionic dimyristoylphosphatidyl glycerol (DMPG) phospholipids. Most of the peptides studied significantly modified the phase transition profile, suggesting the importance of hydrophobic forces for the peptide interactions with the lipid bilayers and their insertion into the bilayer. Indolicidin and tritrpticin are both known to be flexible in aqueous solution, but they adopt turn-turn structures when they bind to and insert in a membrane surface. Pro-to-Ala substitutions in tritrpticin, which result in the formation of a stable α-helix in this peptide, lead to a substantial increase in the peptide interactions with both zwitterionic and anionic phospholipid vesicles. In contrast, the substitution of the three Trp residues by Tyr or Phe resulted in a significant decrease of the peptide's interaction with anionic vesicles and virtually eliminated binding of these peptides to the zwitterionic vesicles. An increase of the cationic charge of the peptide induced much smaller changes to the peptide interaction with all lipid systems than substitution of particular amino acids or modification of the peptide conformation. The presence of multiple lipid domains with a non-uniform peptide distribution was noticed. Slow equilibration of the lipid-peptide systems due to peptide redistribution was observed in some cases. Generally good agreement between the present DSC data and peptide antimicrobial activity data was obtained.  相似文献   

4.
The 13-residue cathelicidins indolicidin and tritrpticin are part of a group of relatively short tryptophan-rich antimicrobial peptides that hold potential as future substitutes for antibiotics. Differential scanning calorimetry (DSC) has been applied here to study the effect of indolicidin and tritrpticin as well as five tritrpticin analogs on the phase transition behaviour of model membranes made up of zwitterionic dimyristoylphosphatidylcholine (DMPC, DMPC/cholesterol) and anionic dimyristoylphosphatidyl glycerol (DMPG) phospholipids. Most of the peptides studied significantly modified the phase transition profile, suggesting the importance of hydrophobic forces for the peptide interactions with the lipid bilayers and their insertion into the bilayer. Indolicidin and tritrpticin are both known to be flexible in aqueous solution, but they adopt turn-turn structures when they bind to and insert in a membrane surface. Pro-to-Ala substitutions in tritrpticin, which result in the formation of a stable alpha-helix in this peptide, lead to a substantial increase in the peptide interactions with both zwitterionic and anionic phospholipid vesicles. In contrast, the substitution of the three Trp residues by Tyr or Phe resulted in a significant decrease of the peptide's interaction with anionic vesicles and virtually eliminated binding of these peptides to the zwitterionic vesicles. An increase of the cationic charge of the peptide induced much smaller changes to the peptide interaction with all lipid systems than substitution of particular amino acids or modification of the peptide conformation. The presence of multiple lipid domains with a non-uniform peptide distribution was noticed. Slow equilibration of the lipid-peptide systems due to peptide redistribution was observed in some cases. Generally good agreement between the present DSC data and peptide antimicrobial activity data was obtained.  相似文献   

5.
Nanoscale structural reorganization of a lipid bilayer membrane induced by a chemical recognition event has been imaged using in situ atomic force microscopy (AFM). Supported lipid bilayers, composed of distearylphosphatidylcholine (DSPC) and a synthetic lipid functionalized with a Cu(2+) receptor, phase-separate into nanoscale domains that are distinguishable by the 9 A height difference between the two molecules. Upon binding of Cu(2+) the electrostatic nature of the receptor changes, causing a dispersion of the receptor molecules and subsequent shrinking of the structural features defined by the receptors in the membrane. Complete reversibility of the process was demonstrated through the removal of metal ions with EDTA.  相似文献   

6.
Kinetics of solid state stability of seven derivatives of 3,5-disubstituted tetrahydro-2H-1,3,5-thiadiazine-2-thione (THTT) of glycine as a model for amino acids and peptide drugs were studied using differential scanning calorimetry (DSC). Each DSC curve for each derivative showed an endothermic peak followed by an exothermic one, which could be attributed to the melting and decomposition, respectively. The decomposition activation energy of each derivative was calculated using the Augis and Bennet, Kissinger equations and Mahadevan approximation. Also, the melting activation energies as well as the thermodynamic parameter (enthalpy) for the investigated derivatives were evaluated. The relative stability of the derivatives in the solid state according to the calculated values of the decomposition activation energy, frequency factors and half-life for each derivative could be determined.  相似文献   

7.
Ice premelting during differential scanning calorimetry   总被引:1,自引:0,他引:1       下载免费PDF全文
PW Wilson  JW Arthur    AD Haymet 《Biophysical journal》1999,77(5):2850-2855
Premelting at the surface of ice crystals is caused by factors such as temperature, radius of curvature, and solute composition. When polycrystalline ice samples are warmed from well below the equilibrium melting point, surface melting may begin at temperatures as low as -15 degrees C. However, it has been reported (. Biophys. J. 65:1853-1865) that when polycrystalline ice was warmed in a differential scanning calorimetry (DSC) pan, melting began at about -50 degrees C, this extreme behavior being attributed to short-range forces. We show that there is no driving force for such premelting, and that for pure water samples in DSC pans curvature effects will cause premelting typically at just a few degrees below the equilibrium melting point. We also show that the rate of warming affects the slope of the DSC baseline and that this might be incorrectly interpreted as an endotherm. The work has consequences for DSC operators who use water as a standard in systems where subfreezing runs are important.  相似文献   

8.
J F Brandts  L N Lin 《Biochemistry》1990,29(29):6927-6940
Data from differential scanning calorimetry (DSC) may be used to estimate very large binding constants that cannot be conveniently measured by more conventional equilibrium techniques. Thermodynamic models have been formulated to describe interacting systems that involve either one thermal transition (protein-ligand) or two thermal transitions (protein-protein) and either 1:1 or higher binding stoichiometry. Methods are described for obtaining binding constants and heats of binding by two different methods: calculation or simulation fitting of data. Extensive DSC data on 2'CMP binding to RNase are presented and analyzed by the two methods. It is found that the methods agree when binding sites are completely saturated, but substantial errors arise in the calculation method when site saturation is incomplete and the transition of liganded molecules overlaps that of unliganded molecules. This arises primarily from an inability to determine TM (i.e., the temperature where concentrations of folded and unfolded protein are equal) under weak-binding conditions. Results from simulation show that the binding constants and heats of binding from the DSC method agree quantitatively with corresponding estimates obtained from equilibrium methods when extrapolated to the same temperature. It was also found from the DSC data that the binding constant decreases with increasing concentration of ligand, which might arise from nonideality effects associated with dimerization of 2'CMP. Simulations show that the DSC method is capable of estimating binding constants for ultratight interactions up to perhaps 10(40) M-1 or higher, while most equilibrium methods fail well below 10(10) M-1. DSC data from the literature on a number of interacting systems (trypsin-soybean trypsin inhibitor, trypsin-ovomucoid, trypsin-pancreatic trypsin inhibitor, chymotrypsin-subtilisin inhibitor, subtilisin BPN-subtilisin inhibitor, RNase S protein-RNase S peptide, avidin-biotin, ovotransferrin-Fe3+, superoxide dismutase-Zn2+, alkaline phosphatase-Zn2+, and assembly of regulatory and catalytic subunits of aspartate transcarbamoylase) were analyzed by simulation fitting or by calculation. Apparent single-site binding constants ranged from ca. 10(5) to 10(20) M-1, while the interaction constant for assembly of aspartate transcarbamoylase was estimated as 10(37) in molarity units. For most of these systems, the DSC interaction constants compared favorably with other literature estimates, for some it did not for reasons unknown, while for still others this represented the first estimate. Simulations show that for proteins having two binding sites for the same ligand within a single cooperative unit, ligand rearrangement will occur spontaneously during a DSC scan as the transition temperature of the unliganded protein is approached.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

9.
Riboflavin binding (or carrier) protein (RfBP) is a monomeric, two-domain protein, originally purified from hens' egg white. RfBP contains nine disulfide bridges; as a result, the protein forms a compact structure and undergoes reversible three-state thermal denaturation. This was demonstrated using a differential scanning calorimetry (DSC) method [Wasylewski M. (2000) J. Prot. Chem. 19(6), 523-528]. It has been shown that the RfBP complex with riboflavin denaturates in a three-state process which may be attributed to sequential unfolding of the RfBP domains. In case of apo RfBP, the ligand binding domain denaturates at a lower temperature than the C-terminal domain. Ligand binding greatly enhances the thermostability of the N-terminal domain, whereas the C-terminal domain thermostability is only slightly affected and, in case of the examined holo RfBPs, the denaturation peaks of both domains merge or cross over. The magnitude of the changes depends on ligand structure. A detailed study of protein concentration effects carried out in this work allowed to estimate not only the thermostability of both domains but also the strength of domain interactions. The DeltaCp, of denaturation was found for C-terminus and N-terminus of RfBP-riboflavin complex to amount to 2.5 and -1.9 kcal mol(-1), respectively. The calculated domain interaction free energy, DeltaGCN, was estimated to be approximately -1580 cal mol(-1) at 67.0 degrees C. This value indicates that the interdomain interaction is of medium strength.  相似文献   

10.
Dynamic analysis of differential scanning calorimetry data   总被引:2,自引:0,他引:2  
The apparent heat capacity function measured by high-sensitivity differential scanning calorimetry contains dynamic components of two different origins: (1) an intrinsic component arising from the finite instrument time response; and (2) a sample component arising from the kinetics of the thermal transition under study. The intrinsic instrumental component is always present and its effect on the shape of the experimental curve depends on the magnitude of the calorimeter response time. Usually, high-sensitivity instruments exhibit characteristic time constants varying from 10 to 100 s. This slow response introduces distortions in the shape of the heat capacity function especially at fast scanning rates. In addition to this instrumental component, dynamic effects due to sample relaxation processes also contribute to the shape of the experimental heat capacity profile. Since the nature and magnitude of these effects are a function of the kinetic parameters of the transition, they can be used to obtain kinetic information. This communication presents a dynamic deconvolution technique directed to remove artificial distortions in the shape of the heat capacity function measured at any scanning rate, and to obtain a kinetic characterization of a thermally induced transition. The kinetic characterization obtained by this method allows the researcher to obtain transition relaxation times as a continuous function of temperature. This technique has been applied to the thermal unfolding of ribonuclease A and the pretransition of dipalmitoylphosphatidylcholine (DPPC). In both systems the transition relaxation times are temperature dependent. For the protein system the relaxation time is very slow below the transition temperature (approximately 30 s) and very fast above Tm (less than 1 s) in agreement with direct kinetic measurements. For the pretransition of DPPC, the relaxation time is maximal at the transition midpoint and of the order of approx. 40 s.  相似文献   

11.
Melting curves of human plasma measured by differential scanning calorimetry (DSC), known as thermograms, have the potential to markedly impact diagnosis of human diseases. A general statistical methodology is developed to analyze and classify DSC thermograms to analyze and classify thermograms. Analysis of an acquired thermogram involves comparison with a database of empirical reference thermograms from clinically characterized diseases. Two parameters, a distance metric, P, and correlation coefficient, r, are combined to produce a 'similarity metric,' ρ, which can be used to classify unknown thermograms into pre-characterized categories. Simulated thermograms known to lie within or fall outside of the 90% quantile range around a median reference are also analyzed. Results verify the utility of the methods and establish the apparent dynamic range of the metric ρ. Methods are then applied to data obtained from a collection of plasma samples from patients clinically diagnosed with SLE (lupus). High correspondence is found between curve shapes and values of the metric ρ. In a final application, an elementary classification rule is implemented to successfully analyze and classify unlabeled thermograms. These methods constitute a set of powerful yet easy to implement tools for quantitative classification, analysis and interpretation of DSC plasma melting curves.  相似文献   

12.
The effect of NaCl on the thermal inactivation of Listeria monocytogenes has been investigated by conventional microbiological techniques and by using differential scanning calorimetry (DSC). Addition of 1.5 M-NaCl to cells grown at lower NaCl concentrations significantly increases the tolerance of cells to mild heat stress (56-62 degrees C). DSC thermograms show five main peaks which are shifted to higher temperatures in the presence of 1.5 M-NaCl. Measurement of loss of viability in the calorimeter gave good correlation between cell death and the first major thermogram peak at two NaCl concentrations. The time course of the loss of this first peak when cells were heated and held at 60 degrees C in the calorimeter matched the loss of viability, whereas the peak attributable to DNA showed little change during this process. The use of DSC to investigate the mechanisms involved in thermal inactivation is discussed.  相似文献   

13.
14.
Interactions between melittin and a variety of negatively-charged lipid bilayers have been investigated by intrinsic fluorescence, fluorescence polarization of 1,6-diphenylhexatriene and differential scanning calorimetry. (1) Intrinsic fluorescence of the single tryptophan residue of melittin shows that binding of this peptide to negatively-charged phospholipids is directly related to the surface charge density, but is unaffected by the physical state of lipids, fluid or gel, single-shell vesicles or unsonicated dispersions. (2) Changes in the thermotropic properties of negatively-charged lipids upon melittin binding allow to differentiate two groups of lipids: (i) A progressive disappearance of the transition, without any shift in temperature, is observed with monoacid C14 lipids such as dimyristoylphosphatidylglycerol and -serine (group 1). (ii) With a second group of lipids (group 2), a transition occurs even at melittin saturation, and two transitions are detected at intermediate melittin content, one corresponding to remaining unperturbed lipids, the other shifted downward by 10–20°C. This second group of lipids is constituted by monoacid C16 lipids, dipalmitoylphosphatidylglycerol and -serine. Phosphatidic acids also enter this classification, but it is the net charge of the phosphate group which allows to discriminate: singly charged phosphatidic acids belong to group 2, whereas totally ionized ones behave like group 1 lipids, whatever the chain length. (3) It is concluded that melittin induces phase separations between unperturbed lipid regions which give a transition at the same temperature as pure lipid, and peptide rich domains in which the stoichiometry is 1 toxin per 8 phospholipids. The properties of such domains depend on the bilayer stability: in the case of C16 aliphatic chains and singly charged polar heads, the lipid-peptide domains have a transition at a lower temperature than the pure lipid. With shorter C14 chains or with two net charges by polar group, the bilayer structure is probably totally disrupted, and the new resulting phase can no longer lead to a cooperative transition.  相似文献   

15.
Thermal denaturation of the B form of double-stranded DNA has been probed by differential scanning calorimetry (DSC) and Raman spectroscopy of 160 base pair (bp) fragments of calf thymus DNA. The DSC results indicate a median melting temperature Tm = 75.5 degrees C with calorimetric enthalpy change delta Hcal = 6.7 kcal/mol (bp), van't Hoff enthalpy change delta HVH = 50.4 kcal/mol (cooperative unit), and calorimetric entropy change delta Scal = 19.3 cal/deg.mol (bp), at the experimental conditions of 55 mg DNA/ml in 5 mM sodium cacodylate at pH 6.4. The average cooperative melting unit (nmelt) comprises 7.5 bp. The Raman signature of 160 bp DNA is highly sensitive to temperature. Analyses of several conformation-sensitive Raman bands indicate the following ranges for thermodynamic parameters of melting: 43 < delta HVH < 61 kcal/mol (cooperative unit), 75 < Tm < 80 degrees C and 6 < (nmelt) < 9 bp, consistent with the DSC results. The changes observed in specific Raman band frequencies and intensities as a function of temperature reveal that thermal denaturation is accompanied by disruption of Watson-Crick base pairs, unstacking of the bases and disordering of the B form backbone. These three types of structural change are highly correlated throughout the investigated temperature range of 20 to 93 degrees C. Raman bands diagnostic of purine and pyrimidine unstacking, conformational rearrangements in the deoxyribose-phosphate moieties, and changes in environment of phosphate groups have been identified. Among these, bands at 834 cm-1 (due to a localized vibration of the phosphodiester group), 1240 cm-1 (thymine ring) and 1668 cm-1 (carbonyl groups of dT, dG and dC), are shown by comparison with DSC results to be the most reliable quantitative indicators of DNA melting. Conversely, the intensities of Raman marker bands at 786 cm-1 (cytosine ring), 1014 cm-1 (deoxyribose ring) and 1092 cm-1 (phosphate group) are largely invariant to melting and are proposed as appropriate standards for intensity normalizations.  相似文献   

16.
何培青  李江  王昉  顾敏芬  沈继红 《生态学报》2009,29(11):5766-5772
采用差示扫描量热法,测定几种南极细菌胞外多糖(简称,EPSs)溶液的结晶、熔融、焓转变以及水合性质等冻结特性,分析了EPSs的浓度和分子量与其抗冻活性的关系.结果表明,在溶液冻结过程中,仅0.25%的Pseudoalteromonas sp.S-15-13 EPSs(分子量,6.2×104Da)可抑制冰核形成,溶液冻结温度较纯水的降低(1.07±0.62)℃;溶液的冻结焓降低说明冰核生长变缓,冰晶形成细小,0.125%的Shewanella sp.5-1-11-4 EPSs(分子量,1.2×103Da)和Moritella sp.2-5-10-1 EPSs(分子量,3.0×103Da)冻结焓分别较纯水的降低17.15%和29.13%,S-15-13 EPSs在0.125%~0.5%的范围内可降低冻结焓,0.125%时冻结焓较纯水的低30%,其不冻水含量为(0.292 ±0.05) g/g.在冰晶熔化过程中,几种EPSs均可降低溶液熔融温度和熔融焓,促进冰晶熔化,使冰晶细小;4.0%的5-1-11-4 EPSs、2-5-10-1 EPSs和0.5% S-15-13 EPSs的熔融温度较纯水的分别降低(2.70±0.15)℃、(2.30±0.39)℃和(4.66±0.42)℃.研究结果阐明EPSs可以通过改变菌体周围水的冻结特性,以抵御冰晶对微生物的损伤,大分子量EPSs对冰晶的抑制作用强于低分子量的.  相似文献   

17.
18.
Thermograms of the exosporium-lacking dormant spores of Bacillus megaterium ATCC 33729, obtained by differential scanning calorimetry, showed three major irreversible endothermic transitions with peaks at 56, 100, and 114 degrees C and a major irreversible exothermic transition with a peak at 119 degrees C. The 114 degrees C transition was identified with coat proteins, and the 56 degrees C transition was identified with heat inactivation. Thermograms of the germinated spores and vegetative cells were much alike, including an endothermic transition attributable to DNA. The ascending part of the main endothermic 100 degrees C transition in the dormant-spore thermograms corresponded to a first-order reaction and was correlated with spore death; i.e., greater than 99.9% of the spores were killed when the transition peak was reached. The maximum death rate of the dormant spores during calorimetry, calculated from separately measured D and z values, occurred at temperatures above the 73 degrees C onset of thermal denaturation and was equivalent to the maximum inactivation rate calculated for the critical target. Most of the spore killing occurred before the release of most of the dipicolinic acid and other intraprotoplast materials. The exothermic 119 degrees C transition was a consequence of the endothermic 100 degrees C transition and probably represented the aggregation of intraprotoplast spore components. Taken together with prior evidence, the results suggest that a crucial protein is the rate-limiting primary target in the heat killing of dormant bacterial spores.  相似文献   

19.
20.
Many methods exist for measuring and studying protein denaturation in vitro. However, measuring protein denaturation in cells under conditions relevant to heat shock presents problems due to cellular complexity and high levels of light scattering that interfere with optical techniques. A general method for measuring protein denaturation in cells using high sensitivity differential scanning calorimetry (DSC) is given. Profiles of specific heat (c(p) vs. temperature) are obtained providing information about transitions in cellular components including the denaturation of proteins. The specific approaches employed with erythrocytes, bacteria, and mammalian cells are described, and an identification of several features of the DSC profiles is given. Protein denaturation on the level of roughly 7-20% occurs for commonly used heat shocks in mammalian cells.  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