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.
A new method has been developed for determining the stability parameters of proteins from their heat-induced transition curves followed by observation of changes in the far-UV circular dichroism (CD). This method of analysis of the thermal denaturation curve of a protein gave values of stability parameters that not only are identical to those measured by the differential scanning calorimetry (DSC), but also are measured with the same error as that observed with a calorimeter. This conclusion has been reached from our studies of the reversible heat-induced denaturation of lysozyme and ribonuclease A at various pH values. For each protein, the conventional method of analysis of the conformational transition curve, which assumes a linear temperature dependence of the pre- and posttransition baselines, gave the estimate of DeltaH(van)(m) (enthalpy change on denaturation at T(m), the midpoint of denaturation) which is significantly lower than DeltaH(cal)(m), the value obtained from DSC measurements. However, if the analysis of the same denaturation curve assumes that a parabolic function describes the temperature dependence of the pre- and posttransition baselines, there exists an excellent agreement between DeltaH(van)(m) and DeltaH(cal)(m) of the protein. The latter analysis is supported by the far-UV CD measurements of the oxidized ribonuclease A as a function of temperature, for the temperature dependence of this optical property of the protein is indeed nonlinear. Furthermore, it has been observed that, for each protein, the constant-pressure heat capacity change (DeltaC(p)) determined from the plots of DeltaH(van)(m) versus T(m) is independent of the method of analysis of the transition curve.  相似文献   

2.
Small monomeric proteins from mesophilic and thermophilic organisms were studied. They have close structural and physical and chemical properties but vary in thermal stability. A thermodynamic analysis of heat unfolding was made and integral enthalpy of unfolding (DeltaH(unf)), heat capacity of hydration (DeltaC(p)(hyd)) and enthalpy of hydration (DeltaH(hyd)) and of the buried surface area (DeltaASA) of nonpolar and polar groups as well as the enthalpy of disruption of intramolecular interaction (DeltaH(int) in gas phase) at 298 K were determined. The absence of correlation between protein thermostability and energetic components suggests that regulatory mechanism of protein thermal stabilization has entropic nature.  相似文献   

3.
We investigated the thermotropic and barotropic bilayer phase behavior of 1-myristoyl-2-oleoyl-sn-glycero-3-phosphocholine (MOPC) and 1-oleoyl-2-myristoyl-sn-glycero-3-phosphocholine (OMPC) by means of the differential scanning calorimetry (DSC) and high-pressure light-transmittance technique. Water could be used as a solvent for measurements at high pressures because of the elevation of the transition temperatures above 0 degrees C by pressurization, whereas aqueous 50 wt.% ethylene glycol solution was used mainly for those at low pressures. Only one phase transition was observed in the DSC thermogram of the MOPC bilayer membrane as an endothermic peak, and also observed at high pressures as an abrupt change of the light-transmittance. The transition was assigned as a main transition between the lamellar gel (L(beta)) and liquid-crystalline (L(alpha)) phases on the basis of the values of enthalpy change (DeltaH) and slope of the transition temperature with respect to pressure (dT/dP). The DSC thermogram of the OMPC bilayer membrane similarly showed a single endothermic peak but two kinds of phase transitions were observed at different temperatures in the light-transmittance profile at high pressures. The extrapolation of the lower-temperature transition in the high-pressure range to an ambient pressure coincided with the transition observed in the DSC thermogram. This transition was identified as a transition between the lamellar crystal (L(c)) and L(alpha) (or L(beta)) phases from the DeltaH and dT/dP values. The higher-temperature transition, appearing only at high pressures, was identified as the L(beta)/L(alpha) transition considering the topological resemblance of its temperature-pressure phase diagram as that of the dioleoylphosphatidylcholine bilayer membrane. The phase diagram of the OMPC bilayer membrane demonstrated that the L(beta) phase cannot exist at pressures below ca. 190 MPa while it can exist stably in a finite temperature range at pressures above the pressure.  相似文献   

4.
The effects of anions on the thermostability of ovotransferrin (oTf) were investigated. The temperature, T(m), causing aggregation of oTf was measured in the presence or absence of anions, and the denaturation temperature, T(m)(DSC), was also determined by differential scanning calorimetry (DSC) in the presence of the citrate anion. We found that some anions (phosphate, sulfate and citrate) raised temperature T(m) of oTf by about 5-7 degrees C. However, neither sodium chloride nor sodium bicarbonate raised T(m) by that much. Temperature T(m) was increased by increasing the concentration of the citrate anion, and was in good agreement with denaturation temperature T(m)(DSC), suggesting that denaturation of the oTf molecules resulted in aggregation of oTf. We also demonstrated that the anions, especially sulfate, repressed the heat-aggregation of liquid egg white.The Van't Hoff plot from the T(m) and DeltaH(d) values revealed that two anion-binding sites were concerned with heat stabilization. These binding sites may have been concerned with sulfate binding (not bicarbonate binding) that is found in the crystal structure of apo-form of oTf, since the bicarbonate anion did not raise T(m).  相似文献   

5.
The formal equilibrium reduction potentials of recombinant electron transport protein, rubredoxin (MW = 7500 Da), from both the mesophilic Clostridium pasteurianum (Topt = 37 degrees C) and hyperthermophilic Pyrococcus furiosus (Topt = 95 degrees C) were recorded as a function of pressure and temperature. Measurements were made utilizing a specially designed stainless steel electrochemical cell that easily maintains pressures between 1 and 600 atm and a temperature-controlled cell that maintains temperatures between 4 and 100 degrees C. The reduction potential of P. furiosus rubredoxin was determined to be 31 mV at 25 degrees C and 1 atm, -93 mV at 95 degrees C and 1 atm, and 44 mV at 25 degrees C and 400 atm. Thus, the reduction potential of P. furiosus rubredoxin obtained under standard conditions is likely to be dramatically different from the reduction potential obtained under its normal operating conditions. Thermodynamic parameters associated with electron transfer were determined for both rubredoxins (for C. pasteurianum, DeltaV degrees = -27 mL/mol, DeltaS degrees = -36 cal K-1 mol-1, and DeltaH degrees = -10 kcal/mol, and for P. furiosus, DeltaV degrees = -31 mL/mol, DeltaS degrees = -41 cal K-1 mol-1, and DeltaH degrees = -13 kcal/mol) from its pressure- and temperature-reduction potential profiles. The thermodynamic parameters for electron transfer (DeltaV degrees, DeltaS degrees, and DeltaH degrees ) for both proteins were very similar, which is not surprising considering their structural similarities and sequence homology. Despite the fact that these two proteins exhibit dramatic differences in thermostability, it appears that structural changes that confer dramatic differences in thermostability do not significantly alter electron transfer reactivity. The experimental changes in reduction potential as a function of pressure and temperature were simulated using a continuum dielectric electrostatic model (DELPHI). A reasonable estimate of the protein dielectric constant (epsilonprotein) of 6 for both rubredoxins was determined from these simulations. A discussion is presented regarding the analysis of electrostatic interaction energies of biomolecules through pressure- and temperature-controlled electrochemical studies.  相似文献   

6.
Structural studies of starches with different water contents   总被引:1,自引:0,他引:1  
The proportion of double helices in starches from a series of pea [rb, rug4-b, rug3-a, and lam-c mutants, and the wild type (WT) parental line], potato and maize (normal and low amylose), and wheat (normal) lines, ranged from about 30-50% on a dry weight basis. In relatively dry starch powders, only about half of the double helices were in crystalline order, this proportion being higher for A-type than for B-type starches. Using starch from WT pea as an example, it was found that increasing water content results in an increase in total crystallinity. When the water content was raised to a level similar to that in excess water, the proportion of crystallinity was close to the proportion of double helices (DH). Measuring crystallinity in starches with a high water content is difficult using traditional methods such as x-ray diffraction. A method was developed, therefore, for determining starch structural characteristics in excess water by measuring the enthalpy of gelatinization transition in quasi-equilibrium differential scanning calorimetry (DSC) experiments. It is suggested that DH% = DeltaH(sp)/DeltaH(DH) x 100%, where DeltaH(sp) and DeltaH(DH) represent the specific enthalpies of gelatinisation transition, DeltaH(sp) being measured as J/g dry starch weight and DeltaH(DH) as J/g DH, in starch. Studies on potato and maize starches in excess water and in 0.6M KCl showed, respectively, that DeltaH(DH) was 36.3 and 35.6 J/g for B-type polymorphs and 33.0 and 35.0 J/g for A-type polymorphs. For C-type starches, such as those from pea, intermediate values of DeltaH(DH), related to the proportions A-/B-polymorphs, should be used. The type of crystallinity in starch can be determined by the shift in peak temperature for thermograms in excess water and in excess 0.6M KCl. For B-polymorphs this shift was found to be approximately 2-3 degrees C and for A-polymorphs approximately 7-12 degrees C. The ratio between ordered areas with both A- and B-polymorphs can be determined from the enthalpies of disruption of each area. These enthalpies can be obtained by deconvolution of bimodal thermograms produced by C-type starches in excess 0.6M KCl. This methodical approach can be applied to all starches that give a sharp gelatinisation thermogram in excess water. Using a range of methods, including DSC, it was found that starch granules from the mutant peas are constructed in a similar way to those from the WT, with B-polymorphs in the centre and A-polymorphs at the periphery of all granules. The proportion of A/B-polymorphs, however, differed between the mutants. It was found that in addition to increasing the total crystallinity, increasing the water content within the granules also resulted in an increase in the proportion of B-polymorphs.  相似文献   

7.
Lipocalin-type prostaglandin (PG) D synthase (L-PGDS) is a dual-functioning protein in the lipocalin family, acting as a PGD(2)-synthesizing enzyme and as an extracellular transporter for small lipophilic molecules. We earlier reported that denaturant-induced unfolding of L-PGDS follows a four-state pathway, including an activity-enhanced state and an inactive intermediate state. In this study, we investigated the thermal unfolding mechanism of L-PGDS by using differential scanning calorimetry (DSC) and CD spectroscopy. DSC measurements revealed that the thermal unfolding of L-PGDS was a completely reversible process at pH 4.0. The DSC curves showed no concentration dependency, demonstrating that the thermal unfolding of L-PGDS involved neither intermolecular interaction nor aggregation. On the basis of a simple two-state unfolding mechanism, the ratio of van't Hoff enthalpy (DeltaH(vH)) to calorimetric enthalpy (DeltaH(cal)) was below 1, indicating the presence of an intermediate state (I) between the native state (N) and unfolded state (U). Then, statistical thermodynamic analyses of a three-state unfolding process were performed. The heat capacity curves fit well with a three-state process; and the estimated transition temperature (T(m)) and enthalpy change (DeltaH(cal)) of the N<-->I and I<-->U transitions were 48.2 degrees C and 190 kJ.mol(-1), and 60.3 degrees C and 144 kJ.mol(-1), respectively. Correspondingly, the thermal unfolding monitored by CD spectroscopy at 200, 235 and 290 nm revealed that L-PGDS unfolded through the intermediate state, where its main chain retained the characteristic beta-sheet structure without side-chain interactions.  相似文献   

8.
《The Journal of cell biology》1993,122(6):1267-1276
There is circumstantial evidence that protein denaturation occurs in cells during heat shock at hyperthermic temperatures and that denatured or damaged protein is the primary inducer of the heat shock response. However, there is no direct evidence regarding the extent of denaturation of normal cellular proteins during heat shock. Differential scanning calorimetry (DSC) is the most direct method of monitoring protein denaturation or unfolding. Due to the fundamental parameter measured, heat flow, DSC can be used to detect and quantitate endothermic transitions in complex structures such as isolated organelles and even intact cells. DSC profiles with common features are obtained for isolated rat hepatocytes, liver homogenate, and Chinese hamster lung V79 fibroblasts. Five main transitions (A-E), several of which are resolvable into subcomponents, are observed with transition temperatures (Tm) of 45-98 degrees C. The onset temperature is approximately 40 degrees C, but some transitions may extend as low as 37-38 degrees C. In addition to acting as the primary signal for heat shock protein synthesis, the inactivation of critical proteins may lead to cell death. Critical target analysis implies that the rate limiting step of cell killing for V79 cells is the inactivation of a protein with Tm = 46 degrees C within the A transition. Isolated microsomal membranes, mitochondria, nuclei, and a cytosolic fraction from rat liver have distinct DSC profiles that contribute to different peaks in the profile for intact hepatocytes. Thus, the DSC profiles for intact cells appears to be the sum of the profiles of all subcellular organelles and components. The presence of endothermic transitions in the isolated organelles is strong evidence that they are due to protein denaturation. Each isolated organelle has an onset for denaturation near 40 degrees C and contains thermolabile proteins denaturing at the predicted Tm (46 degrees C) for the critical target. The extent of denaturation at any temperature can be approximately by the fractional calorimetric enthalpy. After scanning to 45 degrees C at 1 degree C/min and immediately cooling, a relatively mild heat shock, an estimated fraction denaturation of 4-7% is found in hepatocytes, V79 cells, and the isolated organelles other than nuclei, which undergo only 1% denaturation because of the high thermostability of chromatin. Thus, thermolabile proteins appear to be present in all cellular organelles and components, and protein denaturation is widespread and extensive after even mild heat shock.  相似文献   

9.
Crystalline soybean trypsin inhibitor protein undergoes denaturation on heating which is reversed on cooling. In the range of temperature of 35 to 50 degrees C. a solution of the protein consists of a mixture of native and denatured forms in equilibrium with each other. The equilibrium is only slowly established and its final value at any temperature is the same whether a heated, denatured solution of the protein is cooled to the given temperature or whether a fresh solution is raised to that temperature. The kinetics of reversible denaturation of the soybean protein as well as the reversal of denaturation is that of a reversible unimolecular reaction, each process consisting at a given temperature of the same two simultaneous reactions acting in opposite directions. The experimental data on the effect of temperature on the velocity and the equilibrium constants of the opposing reaction were utilized in evaluating the reaction energies and activation energies. The reaction energies for denaturation were found to be as follows:- Change in total heat of reaction DeltaH = 57,000 calories per mole Change in entropy of reaction DeltaS = 180 calories per degree per mole The heat of activation DeltaH(1) (double dagger) for denaturation = 55,000 The heat of activation DeltaH(2) (double dagger) for the reversal of denaturation = -1900 The entropy DeltaS(1) (double dagger) for denaturation = 95 The entropy DeltaS(2) (double dagger) for reversal of denaturation = -84  相似文献   

10.
Differential scanning calorimetry (DSC) and pressure perturbation calorimetry (PPC) were used to characterize thermal phase transitions, membrane packing, and volumetric properties in multilamellar vesicles (MLVs) composed of the polar lipid fraction E (PLFE) isolated from the thermoacidophilic archaeon Sulfolobus acidocaldarius grown at different temperatures. For PLFE MLVs derived from cells grown at 78 degrees C, the first DSC heating scan exhibits an endothermic transition at 46.7 degrees C, a small hump near 60 degrees C, and a broad exothermic transition at 78.5 degrees C, whereas the PPC scan reveals two transitions at approximately 45 degrees C and 60 degrees C. The endothermic peak at 46.7 degrees C is attributed to a lamellar-to-lamellar phase transition and has an unusually low DeltaH (3.5 kJ/mol) and DeltaV/V (0.1%) value, as compared to those for the main phase transitions of saturated diacyl monopolar diester lipids. This result may arise from the restricted trans-gauche conformational changes in the dibiphytanyl chain due to the presence of cyclopentane rings and branched methyl groups and due to the spanning of the lipid molecules over the whole membrane. The exothermic peak at 78.5 degrees C probably corresponds to a lamellar-to-cubic phase transition and exhibits a large and negative DeltaH value (-23.2 kJ/mol), which is uncommon for normal lamellar-to-cubic phospholipid phase transformations. This exothermic transition disappears in the subsequent heating scans and thus may involve a metastable phase, which is irreversible at the scan rate used. Further, there is no distinct peak in the plot of the thermal expansion coefficient alpha versus temperature near 78.5 degrees C, indicating that this lamellar-to-cubic phase transition is not accompanied by any significant volume change. For PLFE MLVs derived from cells grown at 65 degrees C, similar DSC and PPC profiles and thermal history responses were obtained. However, the lower growth temperature yields a higher DeltaV/V ( approximately 0.25%) and DeltaH (14 kJ/mol) value for the lamellar-to-lamellar phase transition measured at the same pH (2.1). A lower growth temperature also generates a less negative temperature dependence of alpha. The changes in DeltaV/V, DeltaH, and the temperature dependence of alpha can be attributed to the decrease in the number of cyclopentane rings in PLFE at the lower growth temperature. The relatively low DeltaV/V and small DeltaH involved in the phase transitions help to explain why PLFE liposomes are remarkably thermally stable and also echo the proposal that PLFE liposomes are generally rigid and tightly packed. These results help us to understand why, despite the occurrence of thermal-induced phase transitions, PLFE liposomes exhibit a remarkably low temperature sensitivity of proton permeation and dye leakage.  相似文献   

11.
The beta form of human cAMP-dependent protein kinase inhibitor (human PKIbeta), a novel heat-stable protein, was isolated with high yield using a bacterial expression system. Assays of PKI activity demonstrated that purified PKIbeta inhibits the catalytic subunit of cAMP-dependent protein kinase. FTIR, Raman spectroscopy and CD experiments implied that human PKIbeta contained only small amounts of alpha-helix and beta-structures, but large amounts of random coil and turn structures, which may explain its high thermostability. The details of its conformational changes in response to heat were studied by CD experiments for the first time, revealing that the protein unfolded at high temperature and refolded when decreased to room temperature.  相似文献   

12.
Carnosine promotes the heat denaturation of glycated protein   总被引:3,自引:0,他引:3  
Glycation alters protein structure and decreases biological activity. Glycated proteins, which accumulate in affected tissue, are reliable markers of disease. Carnosine, which prevents glycation, may also play a role in the disposal of glycated protein. Carnosinylation tags glycated proteins for cell removal. Since thermostability determines cell turnover of proteins, the present study examined carnosine's effect on thermal denaturation of glycated protein using cytosolic aspartate aminotransferase (cAAT). Glycated cAAT (500 microM glyceraldehyde for 72h at 37 degrees C) increased the T(0.5) (temperature at which 50% denaturation occurs) and the Gibbs free energy barrier (DeltaG) for denaturation. The enthalpy of denaturation (DeltaH) for glycated cAAT was also higher than that for unmodified cAAT, suggesting that glycation changes the water accessible surface. Carnosine enhanced the thermal unfolding of glycated cAAT as evidenced by a decreased T(0.5) and a lowered Gibbs free energy barrier. Additionally, carnosine decreased the enthalpy of denaturation, suggesting that carnosine may promote hydration during heat denaturation of glycated protein.  相似文献   

13.
Kozlov AG  Lohman TM 《Biochemistry》2006,45(16):5190-5205
We have previously shown that the linkage of temperature-dependent protonation and DNA base unstacking equilibria contribute significantly to both the negative enthalpy change (DeltaH(obs)) and the negative heat capacity change (DeltaC(p,obs)) for Escherichia coli SSB homotetramer binding to single-stranded (ss) DNA. Using isothermal titration calorimetry we have now examined DeltaH(obs) over a much wider temperature range (5-60 degrees C) and as a function of monovalent salt concentration and type for SSB binding to (dT)(70) under solution conditions that favor the fully wrapped (SSB)(65) complex (monovalent salt concentration >or=0.20 M). Over this wider temperature range we observe a strongly temperature-dependent DeltaC(p,obs). The DeltaH(obs) decreases as temperature increases from 5 to 35 degrees C (DeltaC(p,obs) <0) but then increases at higher temperatures up to 60 degrees C (DeltaC(p,obs) >0). Both salt concentration and anion type have large effects on DeltaH(obs) and DeltaC(p,obs). These observations can be explained by a model in which SSB protein can undergo a temperature- and salt-dependent conformational transition (below 35 degrees C), the midpoint of which shifts to higher temperature (above 35 degrees C) for SSB bound to ssDNA. Anions bind weakly to free SSB, with the preference Br(-) > Cl(-) > F(-), and these anions are then released upon binding ssDNA, affecting both DeltaH(obs) and DeltaC(p,obs). We conclude that the experimentally measured values of DeltaC(p,obs) for SSB binding to ssDNA cannot be explained solely on the basis of changes in accessible surface area (ASA) upon complex formation but rather result from a series of temperature-dependent equilibria (ion binding, protonation, and protein conformational changes) that are coupled to the SSB-ssDNA binding equilibrium. This is also likely true for many other protein-nucleic acid interactions.  相似文献   

14.
Previous studies on the conformation of the monomeric acetylcholinesterase (AChE) from the krait (Bungarus fasciatus) venom showed that the protein possesses a large permanent dipole moment. These studies predicted that thermal irreversible denaturation must occur via partially unfolded states. The thermal stability of Bungarus AChE was determined using capillary electrophoresis (CE) with optimized conditions. Runs performed at convenient temperature scanning rates provided evidence for an irreversible denaturation process according to the Lumry and Eyring model. The mid-transition temperature, T(m), and the effective enthalpy change, DeltaH(m) were determined at different pH. The temperature dependence of the free energy, DeltaG, of Bungarus AChE unfolding was drawn using values of T(m), DeltaH(m) and DeltaC(p) determined by CE. The thermodynamic parameters for the thermal denaturation of the monomeric snake enzyme were compared with those of different dimeric and tetrameric ChEs. It was shown that the changes in the ratio of DeltaH(cal/)DeltaH(vH) and DeltaC(p) reflect the oligomerization state of these proteins. All these results indicate that wild-type monomeric Bungarus AChE is a stable enzyme under standard conditions. However, designed mutants of this enzyme capable of degrading organophosphates have to be engineered to enhance their thermostability.  相似文献   

15.
The first-order nature of the gel-to-liquid crystal phase transition of phospholipid bilayers requires very slow temperature rates in differential scanning calorimetry (DSC) experiments to minimize any rate-dependent distortions. Proportionality of the DSC signal to the rate poses a problem for studies of substrate-supported bilayers that contain very small volumes of the lipid phase. Recently, we described lipid bilayers self-assembled inside nanoporous substrates. The high density of the nanochannels in these structures provides at least a 500-fold increase in the bilayer surface area for the same size of the planar substrate chips. The increased surface area enables the DSC studies. The rate-dependent DSC curves were modeled as a convolution of a conventional van't Hoff shape and a first-order decay curve of the lipid rearrangement. This analysis shows that although confinement of bilayers to the nanopores of approximately 177 nm in diameter results in a more than threefold longer characteristic time of the lipid rearrangement and a decrease in the cooperative unit number, the phase transition temperature is unaffected.  相似文献   

16.
Differential scanning calorimetric (DSC) measurements were performed on the thermal denaturation of lysozyme and lysozyme complexed with N-acetyl-D-glucosamine (GlcNAc) at pH 5.00 (acetate buffer), 4.25 and 2.25 (Gly-HCl buffer). DSC data have been analyzed to obtain denaturation temperature T(d), enthalpy of denaturation DeltaH(D), heat capacity of denaturation DeltaC(pd) and cooperativity index eta. From these thermodynamic parameters, the binding constant K(L) and enthalpy of binding DeltaH(L), for the weak binding of lysozyme with GlcNAc have been determined. The values of K(L) and DeltaH(L) at pH 5.00 and 298 K are 42 +/- 4 M(-1) and -24 +/- 4 kJ mol(-1), respectively, and agree very well with the experimentally determined values from equilibrium and other studies. The binding constant has also been estimated by simulating the DSC curve with varying values of K(L) (T(d)) until it matches the experimental curve.  相似文献   

17.
The conformational and thermal stability of full-length hemagglutinin (HA) of influenza virus (strain X31) has been investigated using a combination of differential scanning calorimetry (DSC), analytical ultracentrifugation, fluorescence, and circular dichroism (CD) spectroscopy as a function of pH. HA sediments as a rosette comprised of 5-6 trimers (31-35 S) over the pH range of 7.4-5.4. The DSC profile of HA in the native state at pH 7.4 is characterized by a single cooperative endotherm with a transition temperature (Tm) of 66 degrees C and unfolding enthalpy (DeltaH(cal)) of 800 kcal x (mol of trimer)(-1). Upon acidification to pH 5.4, there is a significant decrease in the transition temperature (from 66 to 45 degrees C), unfolding enthalpy [from 800 to 260 kcal x (mol of trimer)(-1)], and DeltaH(cal)/DeltaH(vH) ratio (from 3.0 to approximately 1.3). Whereas the far- and near-UV ellipticities are maintained over this pH range, there is an acid-induced increase in surface hydrophobicity and decrease in intrinsic tryptophanyl fluorescence. The major contribution to the DSC endotherm arises from unfolding HA1 domains. The relationship between acid-induced changes in thermal stability and the fusion activity of HA has been examined by evaluating the kinetics and extent of fusion of influenza virus with erythrocytes over the temperature and pH range of the DSC measurements. Surprisingly, X31 influenza virus retains its fusion activity at acidic pH and temperatures significantly below the unfolding transition of HA. This finding is consistent with the notion that the fusion activity of influenza virus may involve structural changes of only a small fraction of HA molecules.  相似文献   

18.
W A Deutschman  F W Dahlquist 《Biochemistry》2001,40(43):13107-13113
The CheY protein isolated from the hyperthermophile Thermotoga maritima is much more resistant to thermally induced unfolding than is its counterpart from the mesophile Bacillus subtilis. To determine the basis of this increased thermostability, the temperature dependence of the free energy of unfolding was determined for these CheY homologues using denaturant-induced unfolding experiments. This allowed comparison of T. maritima CheY with B. subtilis CheY and determination of the thermodynamic qualities responsible for the enhanced thermostability of T. maritima CheY. The stability of the thermophilic CheY protein is a direct result of the increased enthalpy contribution at the temperature of zero entropy, T(s), and the decreased heat capacity change upon unfolding, resulting in a decreased dependence of the free energy of unfolding on temperature. It was found that neither purely entropic nor purely enthalpic contributions alone (as reflected by T(s)) were sufficient to account for the increase in stability.  相似文献   

19.
Branching enzyme (BE) catalyzes formation of the branch points in glycogen and amylopectin by cleavage of the alpha-1,4 linkage and its subsequent transfer to the alpha-1,6 position. We have identified a novel BE encoded by an uncharacterized open reading frame (TK1436) of the hyperthermophilic archaeon Thermococcus kodakaraensis KOD1. TK1436 encodes a conserved protein showing similarity to members of glycoside hydrolase family 57 (GH-57 family). At the C terminus of the TK1436 protein, two copies of a helix-hairpin-helix (HhH) motif were found. TK1436 orthologs are distributed in archaea of the order Thermococcales, cyanobacteria, some actinobacteria, and a few other bacterial species. When recombinant TK1436 protein was incubated with amylose used as the substrate, a product peak was detected by high-performance anion-exchange chromatography, eluting more slowly than the substrate. Isoamylase treatment of the reaction mixture significantly increased the level of short-chain alpha-glucans, indicating that the reaction product contained many alpha-1,6 branching points. The TK1436 protein showed an optimal pH of 7.0, an optimal temperature of 70 degrees C, and thermostability up to 90 degrees C, as determined by the iodine-staining assay. These properties were the same when a protein devoid of HhH motifs (the TK1436DeltaH protein) was used. The average molecular weight of branched glucan after reaction with the TK1436DeltaH protein was over 100 times larger than that of the starting substrate. These results clearly indicate that TK1436 encodes a structurally novel BE belonging to the GH-57 family. Identification of an overlooked BE species provides new insights into glycogen biosynthesis in microorganisms.  相似文献   

20.
The first-order nature of the gel-to-liquid crystal phase transition of phospholipid bilayers requires very slow temperature rates in differential scanning calorimetry (DSC) experiments to minimize any rate-dependent distortions. Proportionality of the DSC signal to the rate poses a problem for studies of substrate-supported bilayers that contain very small volumes of the lipid phase. Recently, we described lipid bilayers self-assembled inside nanoporous substrates. The high density of the nanochannels in these structures provides at least a 500-fold increase in the bilayer surface area for the same size of the planar substrate chips. The increased surface area enables the DSC studies. The rate-dependent DSC curves were modeled as a convolution of a conventional van’t Hoff shape and a first-order decay curve of the lipid rearrangement. This analysis shows that although confinement of bilayers to the nanopores of ∼177 nm in diameter results in a more than threefold longer characteristic time of the lipid rearrangement and a decrease in the cooperative unit number, the phase transition temperature is unaffected.  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