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 406 毫秒
1.
The spin lattice relaxation rates (1T1) of the natural abundance 13C of all seven carbons of α-methyl-d-glucopyranoside were measured in the presence of Mn(II)-concanavalin A. The paramagnetic contribution to the relaxation rates was used to calculate the distance between the Mn(II) site and the saccharide. The results are consistent with the binding of the saccharide in a unique configuration on the surface of the protein with the ?CH2OH (6 carbon) ~9Å, the ?CH3 (7 carbon) 14Å, and carbons 1–5 about 10Å from the Mn(II). Notwithstanding the fact that these distances may be 10% or less in error, these data are in disagreement with a value of greater than 10Å between the saccharide and Mn(II) binding sites determined from X-ray diffraction studies by Edelman et al. ((1972) Proc. Nat. Acad. Sci. USA69, 2580–2584) and Hardman and Ainsworth ((1972) Biochemistry11, 4910–4919).  相似文献   

2.
The hydration properties of Escherichia coli lipids (phosphatidylglycerol, phosphatidylethanolamine) and synthetic 1,2-dioleoyl-sn-glycero-3-phosphocholine in H2O/2H2O mixtures (9:1, v/v) were investigated with 2H-NMR. Comparison of the 2H2O spin lattice relaxation time (T1) as a function of the water content revealed a remarkable quantitative similarity of all three lipid-H2O systems. Two distinct hydration regions could be discerned in the T1 relaxation time profile. (1) A minimum of 11–16 water molecules was needed to form a primary hydration shell, characterized by an average relaxation time of T1 ≈ 90 ms. (2) Additional water was found to be in exchange with the primary hydration shell. The exchange process could be described in terms of a two-site exchange model, assuming rapid exchange between bulk water with T1 = 500 ms and hydration water with T1 = 80–120 ms. Analysis of the linewidth and the residual quadrupole splitting (at low water content) confirmed the size of the primary hydration layer. However, each lipid-water system exhibited a somewhat different linewidth behavior, and a detailed molecular interpretation appeared to be preposterous.  相似文献   

3.
It has been postulated that the heat stabilization of the essential macromolecules in the core of the spore may be produced by dehydration at two levels: (i) the spore is drier at high relative humidity than the vegetative cell and (ii) the core of the spore may be less hydrated than the cortex and the coat. The latter postulate was subjected to experimental testing by 1H-NMR studies of the water signal in the five species of spores and coat and (coat + cortex) preparations. The transverse relaxation rate (1T2) was determined in samples equilibrated at constant relative humidity. To allow for the effect of paramagnetic ions on 1T2 a model system (wool keratin) was studied in the presence of known amounts of Ca(II), Mn(II), Cu(II), Ni(II) and Fe(III). Because of the dominant effect of Mn(II) on 1T2, the effect of small amounts of other metal ions in spores was neglected. The relaxation rate of water at a particular relative humidity and manganese concentration was consistently less for intact spores than for coat or coat + cortex, hence the water in the core is more mobile than in the outer integuments. Sorption isotherm studies have shown that at a particular relative humidity there is about as much water in the core as in the cortex and coat. These two results taken together indicate that the hypothesis that the core is drier than the cortex and coat is incorrect, hence the spore is not heat-stabilized in this way. A theory is proposed in which heat stabilization is attributed to immobilization of essential enzymes and nuclei acids by a solid support, calcium dipicolinate, in a similar fashion to the heat stabilization of enzymes in a charged polymer matrix. It is proposed that stabilization is effected by electrostatic and hydrogen bonds between the calcium dipicolinate and the essential macromolecules. Experiments in progress show that enzymes and DNA are heat-stabilized in vitro by calcium dipicolinate.  相似文献   

4.
Manganese(II)bound at the “tight” metal ion site of unadenylylated glutamine synthetase (E. coli W) has two rapidly exchanging first coordination sphere water molecules. The solvation number was evaluated from a study of the frequency dependence of 1pT1p, the paramagnetic contribution to the longitudinal relaxation rate of solvent protons. The number of rapidly exchanging water molecules is reduced to one in the presence of saturating L-glutamate and to ~0.2 when L-methionine SR-sulfoximine (MSOX) is present. MSOX is a linear competitive inhibitor (KI=3μM) of glutamine synthetase when L-Glu is the substrate. The dissociation constant of MSOX measured by following the 18 fold decrease in 1pT1p (at 48 MHz) is 30μM and is lowered to ~9μM in the presence of ADP. The high affinity of MSOX for the enzyme suggests that this compound mimicks the “transition-state” for the glutamine synthetase reaction. Further evidence for this postulate is found from the dramatic sharpening of the epr spectrum of enzyme-bound Mn(II) in the presence of MSOX and MSOX plus ADP. The intense change in the epr spectrum arises from reduced solvent accessibility to bound Mn(II) and conformational changes produced by binding MSOX and ADP. The suggestion is made from these data that L-Glu and MSOX bind near or directly to the Mn(II) at the “tight” metal ion site in glutamine synthetase isolated from E. coli W.  相似文献   

5.
6.
Molecular motion and molecular organization of human serum low-density lipoprotein (LDL) has been studied in the temperature range ? 30 to 30°C by proton magnetic relaxation. LDL in deuterated Tris-HCl buffer exhibit two mobile phases. The slow-relaxing phase (T1 ? 1.5 s) is assigned to the incompletely deuterated water of the buffer, and the fast-relaxing phase (T1 ? 60 ms) to the fatty acid chains of the lipoprotein core. It has been established that there is a correlation between the state of the outer surface and the interior of the LDL particle: the number of fast-relaxing protons is significantly altered by cooling the system through the freezing point of the buffer or by selecting buffers of different ionic strengths. At room temperature, ~ 30% of the lipid protons of LDL in the 0.1 m buffer and ~ 40% of the lipid protons of LDL in the 0.01 m buffer relax quickly within the time-scale of n.m.r. frequency (24 MHz).  相似文献   

7.
Kinetic studies on the RNase T1-catalyzed transesterification of 12 dinucleoside monophosphates, Np1N2 (N1 = A, C, and U; N2 = A, C, G, and U) at pH 5, 25 °C, and 0.2 m ionic strength, revealed that the catalytic efficiency (kcatKm) for GpN substrates (H. L. Osterman, and F. G. Walz, Jr., 1978, Biochemistry, 17, 4142) was ~106-fold greater than corresponding ApNs and at least 108-fold greater than corresponding CpNs and UpNs. The catalytic activity with ApN substrates survives phenol extraction which indicates (along with other criteria) that it is intrinsic to RNase T1 and is not due to trace contamination by other nucleases. Circumstantial evidence is presented which suggests that homologous GpN and ApN substrates bind productively at different sites on the enzyme. The results of steady-state kinetic studies of RNase T1 with IpNs (N = C and U) were compared with those for GpNs and indicated that the primary effect of the guanine 2-NH2 group is to enhance substrate binding at the primary recognition site by ~2.6 kcal/mol. Values of (kcatKm) showed the order NpC > NpU (N = A, G, and I) which evidences the existence of a subsite for the leaving nucleoside group that prefers cytidine: interactions at this subsite are reflected in kcat rather than Km.  相似文献   

8.
A method is described to measure the oxygen diffusion-concentration product, Do[O2], at any locus that can be probed or labeled using nitroxide radicals. The method is based on the dependence of the spin-lattice relaxation time T1 of the spin label on the bimolecular collision rate with oxygen. Strong Heisenberg exchange between spin label and oxygen contributes directly to T1 of the spin label, while dipolar interactions are negligible. Both time-domain and continuous wave saturation methods for studying T1 are considered. The method has been applied to phospholipid liposomes using fatty acid spin labels. A discontinuity in Do[O2] at the main phase transition was observed.  相似文献   

9.
10.
In the redheaded bunting Emberiza bruniceps, thyroidectomy inhibited premigratory fattening and nocturnal restlessness—two characteristics of avian migration—observed in caged birds during the premigratory period (March/April). Thyroxine (T4) and triiodothyronine (T3) administration in thyroidectomized birds stimulated locomotor activity and restored the loss in body weight. Annual variations in circulating thyroid hormone concentrations revealed a significant rise in T3T4 ratio prior to spring migration in both years studied. This increase in circulating T3T4 ratio may be associated with the development of migratory disposition in this bird. There was no increase in circulating T3T4 ratio prior to autumnal migration, however, plasma T4 increased significantly. Different thyroidal mechanisms are most likely involved in spring and fall migratory periods. While T3 remained low throughout, apart from the characteristic spring rise, high T4 levels in E. bruniceps were associated with periods of reproduction and molting, the latter coinciding partly with autumnal migration.  相似文献   

11.
The effect of cold exposure caused by shearing on serum thyroid hormone (TH) concentrations in sheep kept at an ambient temperature of 8.5°C was studied. While the deep body temperature fell to the lowest level 4 h after shearing the concentration of triiodothyronine (T3) increased to a peak value at that time. Thyroxine (T4) and metabolically inactive reverse triiodothyronine (rT3) levels reached their peak value after 24 h. The T3T4 ratio reached a maximum at about 4 h and rT3T4 and rT3T3 ratios rose to maximum values about 24 h after shearing. This sequence of events suggest a biphasic response to cold—an immediate secretion of TH from the thyroid gland, followed by adaptive alteration in T3 and rT3 generation in the extrathyroidal tissues.  相似文献   

12.
Binding of the chromogenic ligand p-nitrophenyl α-d-mannopyranoside to concanavalin A was studied in a stopped-flow spectrometer. Formation of the protein-ligand complex could be represented as a simple one-step process. No kinetic evidence could be obtained for a ligand-induced change in the conformation of concanavalin A, although the existence of such a conformational change was not excluded. The entire change in absorbance produced on ligand binding occurred in the monophasic process monitored in the stopped-flow spectrometer. The value of the apparent second-order rate constant (ka) for complex formation (ka = 54,000 s?1m? at 25 °C, pH 5.0, Γ/2 0.5) was independent of the protein concentration when the protein was in the range of 233–831 μm in combining sites and in excess of the ligand. The apparent first-order rate constant (k?a) for dissociation of the complex was obtained from the rate constant for the decomposition of the complex upon the addition of excess methyl α-d-mannopyranoside (k?a = 6.2 s?1 at 25 °C, pH 5.0, Γ/2 0.5). The ratio ka?a (0.9 × 104m?1) was in reasonable agreement with value of 1.1 ± 0.1 × 104m?1 determined for the equilibrium constant for complex formation by ultraviolet difference spectrometry. Plots of ln(kaT) and ln(kaT) vs 1T were linear (T is temperature) and were used to evaluate activation parameters. The enthalpies of activation for formation and dissociation of the complex are 9.5 ± 0.3 and 16.8 ± 0.2 kcal/mol, respectively. The unitary entropies of activation for formation and dissociation of the complex are 2.8 ± 1.1 and 1.3 ± 0.7 entropy units, respectively. These entropy changes are much less than those usually associated with substantial changes in the conformation of proteins.  相似文献   

13.
Perturbations induced by melittin on the thermotropism of dimyristoyl-, dipalmitoyl-, distearoylphosphatidylcholine and natural sphingomyelin are investigated and rationalized from data obtained by fluorescence polarization, differential scanning calorimetry and Raman spectroscopy. Depending on the technique and / or experimental conditions used, the observed effects differ at the same lipid to protein molar ratio, due to partial binding of melittin. The binding is more efficient for tetrameric than for monomeric melittin, but in both cases its affinity is weaker for phosphatidylcholine dispersions in the gel phase than for sonicated vesicles. For temperatures T ? Tm efficient binding occurs whatever the initial state of the lipids is. One can summarize the effects induced by melittin on the transition temperature as follows: (i) No upward shift is observed on synthetic phosphatidylcholines when lipid degradation is avoided. This is achieved by using highly purified melittin, phospholipase inhibitors, and / or non-hydrolysable lipids. (ii) Melittin monomer does not change Tm. (iii) When melittin tetramer is stabilized, it decreases Tm by 10–15 deg. C. The transition broadens, and is finally abolished for Ri ? 2. Very similar results are found for natural sphingomyelin. Fluorescence polarization indicates similar changes in order and dynamics of the acyl chains for all lipid studied. For T ? Tm, fluorescence and Raman show that melittin decreases the amount of CH2 groups in ‘trans’ conformation and the intermolecular order of the chains. According to fluorescence data, there is an increase of the rigid-body orientational order at T ? Tm, while from Raman the positional intermolecular order decreases without significant change in the CH2 groups ‘trans’/‘gauche’ ratio.  相似文献   

14.
The α-chymotrypsin-catalyzed hydrolysis rates of p-nitrophenyl cyclopentane-carboxylate (I), p-nitrophenyl indan-2-carboxylate (II), and p-nitrophenyl spiro-[4.4]nonane-2-carboxylate (III) were measured at pH 8.1 in 20% methanol. After correction for variations in reactivity owing to stereoelectronic effects inherent to the substrates, the deacylation rate constants (kc)n of I and II are not significantly different. In (kcKm)n II is 50 times more reactive than I, which demonstrates that the aromatic ring of the former substrate contributes significantly to its reactivity. The nearly equal reactivities of II and III indicate that the enzyme is rather insensitive to the geometry of the nonester-bearing ring of these compounds.  相似文献   

15.
Systematic heat of dilution studies of the self-association of flavin mononucleotide (FMN) have been conducted as a function of ionic strength (0.05 – 2.0 m) and pH (5–9) in aqueous solution. The data are adequately described by the expression QT = ΔH ? (ΔHK)12 (QTcT)12 for an isodesmic self-association. QT is the molar heat of dilution, ΔH and K are the derived enthalpy and equilibrium constants for the process FMN + (FMN)i?1 ? (FMN)i, and cT is the concentration of FMN expressed in monomer units. Typical values derived for the various thermodynamic parameters at 25 °C are ΔG = ?3.56 kcal mol?1, ΔH = ?3.72 kcal mol?1, and ΔS = ?0.54 cal (mol · deg)?1. These data, plus nuclear magnetic resonance evidence (Yagi, K., Ohishi, N., Takai, A., Kawano, K., and Kyogoku, Y., 1976, Biochemistry15, 2877–2880) argue in favor of an open-ended association of flavin molecules. The signs of the various thermodynamic parameters suggest that both hydrophobic and surface energy forces contribute significantly to the association, while the lack of any significant ionic strength dependence indicates the lack of any ionic centers in the association.  相似文献   

16.
A complete titration of phosphatidic acid bilayer membranes was possible for the first time by the introduction of a new anaologue, 1,2-dihexadecyl-sn-glycerol-3-phosphoric acid, which has the advantage of a high chemical stability at extreme pH values. The synthesis of this phosphatidic acid is described and the phase transition behaviour in aqueous dispersions is compared with that of three ester phosphatidic acids; 1,2-dimyristoyl-sn-glycerol-3-phosphoric acid, 1,3-dimyristoylglycerol-2-phosphoric acid and 1,2-dipalmitoyl-sn-glycerol-3-phosphoric acid.The phase transition temperatures (Tt) of aqueous phosphatidic acid dispersions at different degrees of dissociation were measured using fluorescence spectroscopy and 90° light scattering. The Tt values are comparable to the melting points of the solid phosphatidic acids in the fully protonated states, but large differences exist for the charged states.The Tt vs. pH diagrams of the four phosphatidic acids are quite similar and of a characteristic shape. Increasing ionisation results in a maximum value for the transition temperatures at pH 3.5 (pK1). The regions between the first and the second pK of the phosphatidic acids are characterised by only small variations in the transition temperatures (extended plateau) in spite of the large changes occurring in the surface charge of the membranes. The slope of the plateau is very shallow with increasing ionisation. A further decrease in the H+ concentration results in an abrupt change of the transition temperature. The slope of the Tt vs. pH diagram beyond pK2 becomes very steep. This is the  相似文献   

17.
The half-lives of elimination (T12) of 131I-RGG from the body of normal A or Balb/c animals was much longer than the T12 of SJL mice. At all ages, the T12 of normal hybrids (A × SJL, SJL × A, Balb/c × SJL) was similar to or longer than that of the A or Balb/c parents. Thus, in terms of the T12 of normal animals, the SJL responsiveness to 131I-RGG appeared to be a recessive trait. Tolerance could be induced in newborn animals and, in terms of T12, the degree of unresponsiveness at the age of 6 weeks, was the same in A, Balb/c, A × SJL, and Balb/c × SJL animals but was much shorter in SJL mice. Thus, in neonatally induced tolerance, the duration of tolerance was recessive for the SJL type. The average Tbuilt12 after tolerance induction in 3-week-old hybrids (A × SJL, SJL × A, Balb/c × SJL) was similar to that of the A or Balb/c parent, but by the 8th and 12th week it approached the average T12 of the SJL parent. Comparing 8-week-old hybrids, the average T12 was longest in A × SJL hybrids and identical in SJL × A and Balb/c × SJL mice. An examination of T12 distribution in various 8- and 12-week-old crosses and backcrosses revealed a fairly large proportion of individuals with a T12 which was intermediate between SJL and the other parent. There was a tendency for this number to decrease in 12 weeks as compared to 8-week-old mice. In 8-week-old mice, the number of animals with intermediate Tbuilt12 was smallest when SJL was the maternal animal [(SJL × A); SJL × (A × SJL); SJL × (SJL × A)]. There was no link between T12 of tolerant animals and either the immunoglobulin allotype (MuAl/MuA2) or the C5 eniotype (MuB1 positive/MuB1 negative).  相似文献   

18.
To determine the consequences of contact pressure in phyllotaxis, a mathematical model is constructed in which a leaf distribution is represented by a point lattice of n + 1 lattice points at equal intervals on a helix wound around a cylinder. The model is normalized by taking the girth of the cylinder as 1 and by measuring time T in plastochrones, so that n = [T]. r stands for the normalized internode distance (component of the distance between two consecutive lattice points that is parallel to the axis of the cylinder). d stands for the divergence (fraction of a turn between consecutive lattice points). It is assumed that r is a monotonic decreasing function of T such that r(T) → 0 as T → ∞. Contact pressure is represented by the assumption that the minimum geodesic distance between lattice points is maximized. It is shown that if (p, q), with p < q, is the contact phyllotaxis determined when contact pressure first becomes effective, then the continuation of contact pressure requires that the advance to higher phyllotaxis as r decreases must proceed via successive pairs of consecutive terms of the Fibonacci sequence generated by the numbers p and q, namely, p, q, p + q, p + 2q, 2p + 3q, …. The divergence, starting from some value d = 1t + 1a2 + … + 1(an + x) determined by p and q converges to an ideal angle 1t + 1a2 + … + 1an + 1τ, where τ is the golden section. A necessary and sufficient condition for the ideal angle to be 12 + 1τ = τ?2 is that the p and q of the initial contact phyllotaxis be consecutive Fibonacci numbers of the sequence 1, 2, 3, 5, 8, …. It is proved that a sufficient condition for convergence to the ideal angle τ?2 of normal phyllotaxis is that contact pressure begin before T = 5 or before r < 33812 with d initially between 13 and 12.  相似文献   

19.
Using 13C cross-polarization NMR techniques, we have found that the effect of protein on the dynamics of the hydrocarbon interior of a series of biological membranes is to depress the intensity of motion on the nanosecond timescale (i.e., T1 becomes longer) and to enhance the intensity of motion on the timescale of tens of microseconds (i.e., T1p becomes shorter).  相似文献   

20.
Mitochondria isolated from rats chronically fed ethanol demonstrated a marked inability to produce energy. The respiratory control ratio, the ADP/O ratio and state 3 respiration rates were all decreased. Coupled with other data, a progression of ethanol-induced changes is proposed with site I being altered prior to site II. Quantitation of mitochondrial cytochromes revealed decreases in cytochromes b and aa3 and an increase in c1. Evaluation of respiration activity in relation to temperature showed ethanol-induced changes in the transition temperature (Tf) which may have been related to changes in the lipid composition of the inner membrane. Mitochondrial membranes were separated, and analysis of fatty acids and phospholipids was performed. Various fatty acids were altered in both membranes; however, the outer membrane was altered more severely. A decrease in the arachidonate : linoleate ratio was observed only in the outer membrane; however, there was no ethanol-induced change in degree of unsaturation in either membrane. Phospholipid quantitation showed a reduction of total lipid phosphorous/mg protein in both membrane fractions; however, the inner membrane was most affected. Cardiolipin was the only phospholipid in this membrane which remained unaltered. The evidence indicates that the mechanism for ethanol-induced damage to the liver mitochondrion involves lipid compositional changes as well as changes in cytochromes and possibly other proteins.  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