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 31 毫秒
1.
Phenomena occurring in the heme pocket after photolysis of carbonmonoxymyoglobin (MbCO) below about 100 K are investigated using temperature-derivative spectroscopy of the infrared absorption bands of CO. MbCO exists in three conformations (A substrates) that are distinguished by the stretch bands of the bound CO. We establish connections among the A substates and the substates of the photoproduct (B substates) using Fourier-transform infrared spectroscopy together with kinetic experiments on MbCO solution samples at different pH and on orthorhombic crystals. There is no one-to-one mapping between the A and B substates; in some cases, more than one B substate corresponds to a particular A substate. Rebinding is not simply a reversal of dissociation; transitions between B substates occur before rebinding. We measure the nonequilibrium populations of the B substates after photolysis below 25 K and determine the kinetics of B substate transitions leading to equilibrium. Transitions between B substates occur even at 4 K, whereas those between A substates have only been observed above about 160 K. The transitions between the B substates are nonexponential in time, providing evidence for a distribution of substates. The temperature dependence of the B substate transitions implies that they occur mainly by quantum-mechanical tunneling below 10 K. Taken together, the observations suggest that the transitions between the B substates within the same A substate reflect motions of the CO in the heme pocket and not conformational changes. Geminate rebinding of CO to Mb, monitored in the Soret band, depends on pH. Observation of geminate rebinding to the A substates in the infrared indicates that the pH dependence results from a population shift among the substates and not from a change of the rebinding to an individual A substate.  相似文献   

2.
We performed a room temperature molecular dynamics (MD) simulation on a system containing 1 carboxy-myoglobin (MbCO) molecule in a sucrose-water matrix of identical composition (89% [sucrose/(sucrose + water)] w/w) as for a previous trehalose-water-MbCO simulation (Cottone et al., Biophys J 2001;80:931-938). Results show that, as for trehalose, the amplitude of protein atomic mean-square fluctuations, on the nanosecond timescale, is reduced with respect to aqueous solutions also in sucrose. A detailed comparison as a function of residue number evidences mobility differences along the protein backbone, which can be related to a different efficacy in bioprotection. Different heme pocket structures are observed in the 2 systems. The joint distribution of the magnitude of the electric field at the CO oxygen atom and of the angle between the field and the CO unit vector shows a secondary maximum in sucrose, absent in trehalose. This can explain the CO stretching band profile (A substates distribution) differences evidenced by infrared spectroscopy in sucrose- and trehalose-coated MbCO (Giuffrida et al., J Phys Chem B 2004;108:15415-15421), and in particular the appearance of a further substate in sucrose. Analysis of hydrogen bonds at the protein-solvent interface shows that the fraction of water molecules shared between the protein and the sugar is lower in sucrose than in trehalose, in spite of a larger number of water molecules bound to the protein in the former system, thus indicating a lower protein-matrix coupling, as recently observed by Fourier transform infrared (FTIR) experiments (Giuffrida et al., J Phys Chem B 2004;108:15415-15421).  相似文献   

3.
Rebinding and relaxation in the myoglobin pocket   总被引:28,自引:0,他引:28  
The infrared stretching bands of carboxymyoglobin (MbCO) and the rebinding of CO to Mb after photodissociation have been studied in the temperature range 10-300 K in a variety of solvents. Four stretching bands imply that MbCO can exist in four substates, A0-A3. The temperature dependences of the intensities of the four bands yield the relative binding enthalpies and and entropies. The integrated absorbances and pH dependences of the bands permit identification of the substates with the conformations observed in the X-ray data (Kuriyan et al., J. Mol. Biol. 192 (1986) 133). At low pH, A0 is hydrogen-bonded to His E7. The substates A0-A3 interconvert above about 180 K in a 75% glycerol/water solvent and above 270 K in buffered water. No major interconversion is seen at any temperature if MbCO is embedded in a solid polyvinyl alcohol matrix. The dependence of the transition on solvent characteristics is explained as a slaved glass transition. After photodissociation at low temperature the CO is in the heme pocket B. The resulting CO stretching bands which are identified as B substates are blue-shifted from those of the A substates. At 40 K, rebinding after flash photolysis has been studied in the Soret, the near-infrared, and the integrated A and B substates. All data lie on the same rebinding curve and demonstrate that rebinding is nonexponential in time from at least 100 ns to 100 ks. No evidence for discrete exponentials is found. Flash photolysis with monitoring in the infrared region shows four different pathways within the pocket B to the bound substates Ai. Rebinding in each of the four pathways B----A is nonexponential in time to at least 10 ks and the four pathways have different kinetics below 180 K. From the time and temperature dependence of the rebinding, activation enthalpy distributions g(HBA) and preexponentials ABA are extracted. No pumping from one A substate to another, or one B substate to another, is observed below the transition temperature of about 180 K. If MbCO is exposed to intense white light for 10-10(3) s before being fully photolyzed by a laser flash, the amplitude of the long-lived states increases. The effect is explained in terms of a hierarchy of substates and substate symmetry breaking. The characteristics of the CO stretching bands and of the rebinding processes in the heme pocket depend strongly on the external parameters of solvent, pH and pressure. This sensitivity suggests possible control mechanisms for protein reactions.  相似文献   

4.
Two-dimensional (2D) infrared vibrational echoes were performed on horse heart carbonmonoxymyoglobin (MbCO) in water over a range of temperatures. The A(1) and A(3) conformational substates of MbCO are found to have different dephasing rates with different temperature dependences. A frequency-frequency correlation function derived from molecular dynamics simulations on MbCO at 298 K is used to calculate the vibrational echo decay. The calculated decay shows substantial agreement with the experimentally measured decays. The 2D vibrational echo probes protein dynamics and provides an observable that can be used to test structural assignments for the MbCO conformational substates.  相似文献   

5.
Embedding protein in sugar systems of low water content enables one to investigate the protein dynamic-structure function in matrixes whose rigidity is modulated by varying the content of residual water. Accordingly, studying the dynamics and structure thermal evolution of a protein in sugar systems of different hydration constitutes a tool for disentangling solvent rigidity from temperature effects. Furthermore, studies performed using different sugars may give information on how the detailed composition of the surrounding solvent affects the internal protein dynamics and structural evolution. In this work, we compare Fourier transform infrared spectroscopy measurements (300-20 K) on MbCO embedded in trehalose, sucrose, maltose, raffinose, and glucose matrixes of different water content. At all the water contents investigated, the protein-solvent coupling was tighter in trehalose than in the other sugars, thus suggesting a molecular basis for the trehalose peculiarity. These results are in line with the observation that protein-matrix phase separation takes place in lysozyme-lactose, whereas it is absent in lysozyme-trehalose systems; indeed, these behaviors may respectively be due to the lack or presence of suitable water-mediated hydrogen-bond networks, which match the protein surface to the surroundings. The above processes might be at the basis of pattern recognition in crowded living systems; indeed, hydration shells structural and dynamic matching is first needed for successful come together of interacting biomolecules.  相似文献   

6.
AIMS: The efficiency of trehalose, sucrose and maltose to protect Lactobacillus bulgaricus during drying has been evaluated in bacteria grown at low water activity. METHODS AND RESULTS: Bacteria were grown in MRS (control), and in MRS supplemented with sucrose (MRS-sucrose) or with polyethyleneglycol (PEG) (MRS-PEG) as low water activity media. The growth in low water activity media (MRS-sucrose and MRS-PEG) prior to drying enhanced the effectiveness of trehalose as thermoprotectant during drying. The efficiency of sucrose was improved when bacteria were grown in MRS-sucrose. On the other hand, the growth in both low water activity media did not affect the efficiency of maltose. The damage produced during dehydration has been evaluated by means of growth kinetics in milk. The preservation of bacteria dehydrated with sucrose, after growing them in MRS-sucrose, appears to be as efficient as the dehydration with trehalose. CONCLUSIONS: The growth of L. bulgaricus in low water activity media enhances the protective action of trehalose and sucrose. SIGNIFICANCE AND IMPACT OF THE THE STUDY: These results may aid the dairy industry to improve the recovery of the starters at low costs after preservation processes.  相似文献   

7.
The maximum entropy method (MEM) is used to numerically invert the kinetics of ligand rebinding at low temperatures to obtain the underlying two-dimensional distribution of activation enthalpies and entropies, g(H,S). A global analysis of the rebinding of carbon monoxide (CO) to myoglobin (Mb), monitored in the Soret band at temperatures from 60 to 150 K, is performed using a Newton-Raphson optimization algorithm. The MEM approach describes the data much better than traditional least-squares analyses, reducing chi 2 by an order of magnitude. The MEM resolves two barrier distributions suggestive of rebinding to different bound conformations of MbCO, the so-called A1 and A3 substates, whose activation barriers have been independently estimated from kinetics monitored in the infrared. The distribution corresponding to A3 possesses higher activation entropies, also consistent with infrared measurements. Within an A substate, correlations of S and H are recovered qualitatively from simulated data but can be difficult to obtain from experimental data. When the rebinding measured at 60 K is excluded from the inversion, two peaks are no longer clearly resolved. Thus, data of very high quality are required to unambiguously determine the kinetic resolvability of subpopulations and the shape of the barrier distribution for a single A substate.  相似文献   

8.
The ability to cryopreserve a stage of Anopheles mosquitoes would facilitate the development of strains incapable of transmitting malaria. Cryopreservation requires that the freezable water in cell systems be removed or rendered incapable of undergoing ice formation. The present study was concerned with the rate at which water is removed from lst instar larvae of Anopheles gambiae by air-drying, with the extent of dehydration that the larvae will tolerate, and with the effect of trehalose and sucrose on both drying kinetics and survival. Eighty-one percent of the larvae are water. Air-drying removes 90% of that water in approximately 20 min. Survivals after partial dehydration are highest if the larvae are rehydrated in 1/2x isotonic saline (0.13 osm); they are poorest if rehydrated in water or 0.13 osm sucrose. In the former, about 34% survive the removal of half the water, but next to none survive the loss of >70% initial water. Prior exposure to 0.2 M trehalose for as little as 1 min slows the drying rate and increases the tolerance of the larvae to dehydration. With 30-min exposure, 88% survive the loss of 50% of their water and 63% survive the loss of 75%. Protection is abolished with 0.4 M trehalose. The results are similar with sucrose. It is substantially reduced if sugar-exposed larvae are briefly washed with water prior to drying. The protection appears not to be related to the decreased drying rate. Rather it appears related, by an unknown mechanism, to the presence of sugar on the outer surface of the larvae.  相似文献   

9.
Trehalose is the most effective carbohydrate in preserving the structure and function of biological systems during dehydration and subsequent storage. We have studied the kinetics of protein inactivation in amorphous glucose/sucrose (1:10, w/w) and glucose/trehalose (1:10, w/w) systems, and examined the relationship between protein preservation, phase separation and crystallization during dry storage. The glucose/trehalose system preserved glucose-6-phosphate dehydrogenase better than did the glucose/sucrose system with the same glass transition temperature (Tg). The Williams-Landel-Ferry kinetic analysis indicated that the superiority of the glucose/trehalose system over the glucose/sucrose system was possibly associated with a low free volume and a low free volume expansion at temperatures above the Tg. Phase separation and crystallization during storage were studied using differential scanning calorimetry, and three separate domains were identified in stored samples (i.e., sugar crystals, glucose-rich and disaccharide-rich amorphous domains). Phase separation and crystallization were significantly retarded in the glucose/trehalose system. Our data suggest that the superior stability of the trehalose system is associated with several properties of the trehalose glass, including low free volume, restricted molecular mobility and the ability to resist phase separation and crystallization during storage.  相似文献   

10.
The purpose of this study was to correlate the effectiveness of the lysoPC to disrupt bilayers with the effects of trehalose and sucrose on the hydration sites of a lipid bilayer. The vibration frequencies of carbonyls and phosphates was measured at 18 degrees C for different ratios of monomyristoylphosphatidylcholine and dimyristoylphosphatidylcholine vesicles prepared in water, sucrose and trehalose. The disruption point of the bilayer, evaluated by following the changes in the turbidity of the suspension of unilamellar vesicles, was decreased when the vesicles were prepared in 100 mM sucrose. The increase of the lytic action is directly related to the extent of hydration of the carbonyl populations. It is interpreted that the insertion of the sucrose molecule in the interface causes local changes in interfacial structure, such as the dehydration of the second population of the carbonyls that may be identified as defects of packing. In contrast, the insertion of trehalose by replacing water simultaneously at the carbonyls and the phosphates does not cause defects of packing. For this reason, the lytic action is produced at a concentration very similar to that found in water.  相似文献   

11.
Conformational substates of B-DNA had been observed so far in synthetic oligonucleotides but not in naturally occurring highly polymeric B-DNA. Our low-temperature experiments show that native B-DNA from salmon testes and the d(CGCGAATTCGCG)2 dodecamer have the same BI and BII substates. Nonequilibrium distribution of conformer population was generated by quenching hydrated unoriented films to 200 K, and isothermal structural relaxation toward equilibrium by interconversion of substates was followed by Fourier transform infrared spectroscopy. BI interconverts into BII on isothermal relaxation at 200 K, whereas on slow cooling from ambient temperature, BII interconverts into BI. Our estimation of the dodecamer's BI-to-BII conformer substate population by curve resolution of the symmetrical stretching vibration of the ionic phosphate is 2.4 +/- 0.5 to 1 at 200 K, and it is 1.3 +/- 0.5 to 1 between 270 and 290 K. Pronounced spectral changes upon BI-to-BII interconversion are consistent with base destacking coupled with migration of water from ionic phosphate toward the phosphodiester and sugar moieties. Nonspecific interaction of proteins with the DNA backbone could become specific by induced-fit-type interactions with either BI or BII backbone conformations. This suggests that the BI-to-BII substate interconversion could be a major contributor to the protein recognition process.  相似文献   

12.
The nature of the interaction responsible for the inhibition of protein unfolding and subsequent damage by sugars during dehydration is unclear. The relationship between sample moisture content measured by coulometric Karl Fischer titration and the apparent moisture content predicted by the area of the protein side chain carboxylate band at approximately 1580 cm-1 in infrared spectra of dried protein-sugar samples was examined. For samples in which a high level of native protein structure was retained in the dried solid, the apparent moisture content predicted by the carboxylate band area was greater than the actual moisture content, indicating that protection results from direct sugar-protein hydrogen bonding and not entrapment of water at the protein surface. Further, we show that the degree of structural protection conferred by sucrose and trehalose apparent in second derivative, amide I infrared spectra, correlates with the extent of hydrogen bonding between sugar and protein. The failure of dextran to inhibit dehydration-induced lysozyme unfolding is shown to result from the inability of the polymer to hydrogen bond adequately to the protein. Therefore, formation of an amorphous phase alone is not sufficient to maintain protein structure during dehydration. Glucose hydrogen bonds to a high degree with dried lysozyme, but is incapable of inhibiting lyophilization-induced protein unfolding in the absence of an effective cryoprotectant. However, the addition of polyethylene glycol, which is known to protect proteins during freezing, but not drying, to glucose protected lysozyme structure during lyophilization. Together, these results show that hydrogen bonding between carbohydrate and protein is necessary to prevent dehydration-induced protein damage. However, hydrogen bonding alone is not sufficient to protect proteins during lyophilization in the absence of adequate freezing protection.  相似文献   

13.
We report on room temperature electron transfer in the reaction center (RC) complex purified from Rhodobacter sphaeroides. The protein was embedded in trehalose-water systems of different trehalose/water ratios. This enabled us to get new insights on the relationship between RC conformational dynamics and long-range electron transfer. In particular, we measured the kinetics of electron transfer from the primary reduced quinone acceptor (Q(A)(-)) to the primary photo oxidized donor (P(+)), by time-resolved absorption spectroscopy, as a function of the matrix composition. The composition was evaluated either by weighing (liquid samples) or by near infrared spectroscopy (highly viscous or solid glasses). Deconvolution of the observed, nonexponential kinetics required a continuous spectrum of rate constants. The average rate constant ( = 8.7 s(-1) in a 28% (w/w) trehalose solution) increases smoothly by increasing the trehalose/water ratio. In solid glasses, at trehalose/water ratios > or = 97%, an abrupt increase is observed ( = 26.6 s(-1) in the driest solid sample). A dramatic broadening of the rate distribution function parallels the above sudden increase. Both effects fully revert upon rehydration of the glass. We compared the kinetics observed at room temperature in extensively dried water-trehalose matrices with the ones measured in glycerol-water mixtures at cryogenic temperatures and conclude that, in solid trehalose-water glasses, the thermal fluctuations among conformational substates are inhibited. This was inferred from the large broadening of the rate constant distribution for electron transfer obtained in solid glasses, which was due to the free energy distribution barriers having become quasi static. Accordingly, the RC relaxation from dark-adapted to light-adapted conformation, which follows primary charge separation at room temperature, is progressively hindered over the time scale of P(+)Q(A)(-) charge recombination, upon decreasing the water content. In solid trehalose-water glasses the electron transfer process resulted much more affected than in RC dried in the absence of sugar. This indicated a larger hindering of the internal dynamics in trehalose-coated RC, notwithstanding the larger amount of residual water present in comparison with samples dried in the absence of sugar.  相似文献   

14.
Larvae of the Antarctic midge, Belgica antarctica (Diptera: Chironomidae), are frequently exposed to dehydrating conditions on the Antarctic Peninsula. In this study, we examined how rates and levels of dehydration alter heat and cold tolerance and how these relate to levels of trehalose within the insect. When dehydrated, larvae tolerated cold and heat stress more effectively, although resistance to cold was more pronounced than heat resistance. Slow dehydration was more effective than rapid dehydration in increasing temperature tolerance. Severe dehydration (50% reduction in water content) caused a much greater increase in temperature tolerance than did mild dehydration (e.g. 10% water loss). Larvae severely dehydrated at a slow rate (98% RH) were more temperature tolerant than those dehydrated quickly (0 or 75% RH). These results indicate that the slower dehydration rate allows the larvae to more effectively respond to reduced water levels and that physiological adjustments to desiccation provide cross tolerance to cold and heat. Levels of trehalose increased during dehydration and are likely a major factor increasing subsequent cold and heat resistance. This hypothesis was also supported by experimental results showing that injection of trehalose enhanced resistance to temperature stress and dehydration. We conclude that changes in temperature tolerance in B. antarctica are linked to the rate and severity of dehydration and that trehalose elevation is a probable mechanism enhancing this form of cross tolerance.  相似文献   

15.
The kinetic properties of the three taxonomic A substates of sperm whale carbonmonoxy myoglobin in 75% glycerol/buffer are studied by flash photolysis with monitoring in the infrared stretch bands of bound CO at nu(A0) approximately 1967 cm-1, nu(A1) approximately 1947 cm-1, and nu(A3) approximately 1929 cm-1 between 60 and 300 K. Below 160 K the photodissociated CO rebinds from the heme pocket, no interconversion among the A substates is observed, and rebinding in each A substate is nonexponential in time and described by a different temperature-independent distribution of enthalpy barriers with a different preexponential. Measurements in the electronic bands, e.g., the Soret, contain contributions of all three A substates and can, therefore, be only approximately modeled with a single enthalpy distribution and a single preexponential. The bond formation step at the heme is fastest for the A0 substate, intermediate for the A1 substate, and slowest for A3. Rebinding between 200 and 300 K displays several processes, including geminate rebinding, rebinding after ligand escape to the solvent, and interconversion among the A substates. Different kinetics are measured in each of the A bands for times shorter than the characteristic time of fluctuations among the A substates. At longer times, fluctuational averaging yields the same kinetics in all three A substates. The interconversion rates between A1 and A3 are determined from the time when the scaled kinetic traces of the two substates merge. Fluctuations between A1 and A3 are much faster than those between A0 and either A1 or A3, so A1 and A3 appear as one kinetic species in the exchange with A0. The maximum-entropy method is used to extract the distribution of rate coefficients for the interconversion process A0 <--> A1 + A3 from the flash photolysis data. The temperature dependencies of the A substate interconversion processes are fitted with a non-Arrhenius expression similar to that used to describe relaxation processes in glasses. At 300 K the interconversion time for A0 <--> A1 + A3 is 10 microseconds, and extrapolation yields approximately 1 ns for A1 <--> A3. The pronounced kinetic differences imply different structural rearrangements. Crystallographic data support this conclusion: They show that formation of the A0 substate involves a major change of the protein structure; the distal histidine rotates about the C(alpha)-C(beta) bond, and its imidazole sidechain swings out of the heme pocket into the solvent, whereas it remains in the heme pocket in the A1 <--> A3 interconversion. The fast A1 <--> A3 exchange is inconsistent with structural models that involve differences in the protonation between A1 and A3.  相似文献   

16.
The kinetics of oxygen geminate binding with the taxonomic substates of MbO2 are reported. The maximum entropy method was used to analyze the rebinding kinetics of MbCO and MbO2 monitored in the Soret. The resulting rate distributions were found to consist of a small number of overlapping bands. A global parametric fit of a series of rate distributions recorded at several temperatures was performed using a Gaussian basis set to resolve the individual enthalpy distributions P(H). This approach was first validated by showing that the well-documented taxonomic substates of MbCO could be recovered. The method was then applied to MbO2. Three taxonomic substates were identified at pH 4.8, whereas only two of them contribute to oxygen geminate rebinding at pH 7.0. These findings show that, similarly to MbCO, MbO2 also exists as three photolyzable and kinetically different taxonomic substates and suggest reconsidering the issue of the photolysis quantum yield of MbO2.  相似文献   

17.
Disaccharides such as sucrose and trehalose play an important role in stabilizing cellular structures during dehydration. In fact, most organisms that are able to survive desiccation accumulate high concentrations of sugars in their cells. The mechanisms involved in the stabilization of cellular membranes in the dry state have been investigated using model membranes, such as phosphatidylcholine liposomes. It has been proposed that the lyoprotection of liposomes depends on the depression of the gel to liquid-crystalline phase transition temperature (T(m)) of the dry membranes below ambient and on the prevention of membrane fusion by sugar glass formation, because both lead to leakage of soluble content from the liposomes. Since fusion is prevented at lower sugar/lipid mass ratios than leakage, it has been assumed that more sugar is needed to depress T(m) than to prevent fusion. Here, we show that this is not the case. In air-dried egg phosphatidylcholine liposomes, T(m) is depressed by >60 degrees C at sucrose/lipid mass ratios 10-fold lower than those needed to depress fusion to below 20%. In fact, T(m) is significantly reduced at mass ratios where no bulk sugar glass phase is detectable by Fourier transform infrared spectroscopy or differential scanning calorimetry. A detailed analysis of the interactions of sucrose with the P=O, C=O, and choline groups of the lipid and a comparison to published data on water binding to phospholipids suggests that T(m) is reduced by sucrose through a "water replacement" mechanism. However, the sucrose/lipid mass ratios necessary to prevent leakage exceed those necessary to prevent both phase transitions and membrane fusion. We hypothesize that kinetic phenomena during dehydration and rehydration may be responsible for this discrepancy.  相似文献   

18.
MutT distinguishes substrate 8-oxo-dGTP from dGTP and also 8-oxo-dGMP from dGMP despite small differences of chemical structures between them. In this paper we show by the method of molecular dynamics simulation that the transition between conformational substates of MutT is a key mechanism for a high-resolution molecular recognition of the differences between the very similar chemical compounds. (1) The native state MutT has two conformational substates with similar free energies, each characterized by either open or closed of two loops surrounding the substrate binding active site. Between the two substates, the open substate is more stable in free MutT and in dGMP-MutT complex, and the closed substate is more stable in 8-oxo-dGMP-MutT complex. (2) Conformational fluctuation of the open substate is much larger than that of the closed substate. An estimate of associated entropy difference was found to be consistent with the experimentally found difference of entropy contribution to the binding free energies of the two molecules. (3) A hydrogen bond between H7 atom of 8-oxo-dGMP and the sidechain of Asn119 plays a crucial role for maintaining the closed substate in 8-oxo-dGMP-MutT complex. When this hydrogen bond is absent in the H7-deficient dGMP-MutT complex, the closed substate is no more maintained and transition to the more entropically-favored open substate is induced. (4) Thus, this mechanism of the hydrogen bond controlling the relative stabilities of the drastically different two conformational substates enhances the resolution to recognize the small difference of the chemical structures between the two molecules, dGMP and 8-oxo-dGMP.  相似文献   

19.
AIM: This work determines the efficiency of trehalose on the preservation by heat or osmotic drying of a strain of Lactobacillus delbrueckii ssp. bulgaricus. Cell recovery at different trehalose concentrations during drying correlated with the surface properties and osmotic response of cells after rehydration. METHODS AND RESULTS: Bacteria were dried in the presence of glycerol, trehalose, sucrose at 70 degrees C and at 20 degrees C. Trehalose attenuates the loss of viability at 0.25 m. At this concentration, the osmotic response and zeta potential of the bacteria were comparable with the nondried ones. CONCLUSIONS: Trehalose diminishes significantly the damage produced by dehydration both when the bacteria are dried by heating or subjected to osmotic dehydration. This effect appears related to the preservation of the permeability to water and the surface potential of the bacteria. SIGNIFICANCE AND IMPACT OF THE STUDY: Dehydration occurring during heating or during osmosis appears to have similar effects. As dehydration-induced damage is in correlation with osmotic response recovery and is hindered or buffered by the presence of trehalose, it may be related to water eliminated from biological structures involved in water permeation.  相似文献   

20.
The spectroscopy of horseradish peroxidase with and without the substrate analogue benzohydroxamic acid (BHA) was monitored in different solvents as a function of the temperature in the interval from 10 to 300 K. Thermal broadening of the Q(0,0) optical absorption band arises mainly from interaction of the electronic pi --> pi transition with the heme vibrations. In contrast, the width of the IR absorption band of CO bound to heme is controlled by the coupling of the CO transition moment to the electric field of the protein matrix. The IR bandwidth of the substrate free enzyme in the glycerol/H2O solvent hardly changes in the glassy matrix and strongly increases upon heating above the glass transition. Heating of the same enzyme in the trehalose/H2O glass considerably broadens the band. The binding of the substrate strongly diminishes the temperature broadening of the CO band. This result is consistent with the view that the BHA strongly reduces the amplitude of vibrations of the heme pocket environment. Unusually strong thermal broadening of the CO band above the glass transition is interpreted to be caused by thermal population of a very flexible excited conformational substate. The thermal broadening of the same band in the trehalose glass is caused by an increase of the protein vibrational amplitude in each of the conformational substates, their population being independent of the temperature in the glassy matrix.  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