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 62 毫秒
1.
A detailed 220-MHz NMR study has been made of the conformational properties for the homodinucleotide adenylyl-3' leads to 5'-adenosine, ApA, in D2O. Unambiguous signal assignments of all proton signals were made with the aid of selectively deuterated nucleotidyl units, ApA, ApA, and D-8ApA, and complete, accurate sets of NMR parameters were derived by simulation-iteration methods. Sets of limiting chemical shifts and coupling values were also obtained for ApA and constituent monomers 3'-AMP and 5'-AMP at infinite dilution and at identical ionization states for assessment of dimerization effects. Conformational properties were evaluated quantitatively for most of the conformational bonds of ApA and these are consistent with two compact folded dynamically averaged structures, a base-stacked right helical structure, I, characterized as anti, C3'-endo, g-, w,w' (320,330 degrees), g'g', gg, C3'-endo, anti, and a more loosely base-stacked loop structure, II, with anti, C3'-endo, g-, w,w' (80 degrees, 50 degrees), g'g', gg, C3'-endo, anti orientations. Dimerization produces a number of nucleotidyl conformational changes including a shift in ribose equilibrium C2'-endo (S) in equilibrium C3'-endo (N) in favor of C3'-endo in both Ap- and -pA (60:40 vs. 35:65 in monomers), a change in glycosidic torsion angle chiCN toward 0 degrees, and a greater locking-in of rotamers along bonds involved in the phosphodiester backbone. Moreover, there is clear evidence that the transitions from S leads to N forms and chiCN leads to 0 degrees are directly related to base stacking in ApA. Finally, ApA exists in solution as an equilibrium between I, II and an unstacked form(s) with as yet undetermined conformational features. Since C4'-C5', C5'-O5', and C3'-O3' bonds possess exceptional conformational stabilities, it is proposed that destacking occurs primarily by rotation about P-O5' and/or O3'-P. Predominant factors influencing the overall ApA conformation are thus base-base interaction and flexibility about P-O5' and O3'-P, with change of ribose conformation occurring in consequence of an alteration of chiCN, the latter in turn being governed by the need for maximum eta overlap of stacked adenine rings.  相似文献   

2.
The Bacillus subtilis catabolite repression HPr (Crh) exhibits 45% sequence identity when compared to histidine-containing protein (HPr), a phosphocarrier protein of the phosphoenolpyruvate:carbohydrate phosphotransferase system. We report here that Crh preparations contain a mixture of monomers and homodimers, whereas HPr is known to be monomeric in solution. The dissociation rate of dimers is very slow (t1/2 of about 10 hours), and the percentage of dimers in Crh preparations increases with rising temperature or protein concentration. However, at temperatures above 25 degrees C and a protein concentration of 10 mg/ml, Crh dimers slowly aggregate. Typically, NMR spectra recorded at 25 degrees C showed the coexistence of both forms of Crh, while in Crh solutions kept at 35 degrees C, almost exclusively Crh monomers could be detected. Circular dichroism analysis revealed that the monomeric and dimeric forms of Crh are well folded and exhibit the same overall structure. The physiological significance of the slow Crh monomer/dimer equilibrium remains enigmatic.  相似文献   

3.
Formyltetrahydrofolate synthetase monomers are converted to catalytically active tetramers in the presence of monovalent cations. The stoichiometry of the reaction is 4M + 2C+ in equilibrium M4C2(2+). A positive deltaS compensates for an unfavorable positive deltaH so that the overall reaction is exergonic. Both deltaH and deltaS become more positive as the temperature is increased. Association of subunits of the enzyme prepared from Clostridium cylindrosporum is second order with respect to monomer concentration, consistent with a rate-determining dimerization step. Activation parameters for this step at 20 degrees are: deltaG, 12.6 kcal mol-1; deltaH, 12.5 kcal mol-1; deltaS, -05 e.u. The rate-limiting step for the cation-dependent association of Clostridium acidi-urici monomers is believed to be a conformational alteration since first order kinetics is observed. The Eyring plot of the kinetic data obtained for the C. acidi-urici system has a sharp break at 15 degrees. Activation parameters for cation-induced association at 20 degrees are: deltaG, 21.5 kcal mol-1; deltaH, 14.0 kcal mol-1; deltaS, -26.6 e.u.  相似文献   

4.
The conformational properties of GpU modified with the reactive derivative of benzo[a]pyrene, (+/-)-7beta,8alpha-dihydroxy-9alpha,10alpha-epoxy-7,8,9,10-tetrahydrobenzo[a]pyrene, has been investigated utilizing circular dichroism spectroscopy. Binding of this carcinogen to the N2 of G residues in GpU resulted in the formation of four compounds (I to IV) representing two pairs of diastereoisomers. The molar ellipticity values of the modified dimers were approximately twofold higher than those of the modified guanosine monomers. These values were decreased appreciably when the spectra of the dimers were obtained at 80 degrees C or in methanol rather than at 25 degrees C in water, suggesting that under the latter conditions there is a stacking interaction between the carcinogen and the neighboring uridine residue. Based on these results, a conformation is proposed for modified GpU. It includes insertion of the benzo[a]pyrene moiety, by rotation of the modified guanine residue about its glycoside bond, coplanar to the neighboring uridine and perpendicular to the phosphodiester backbone.  相似文献   

5.
Actin polymerization. The mechanism of action of cytochalasin D   总被引:16,自引:0,他引:16  
Fluorescence changes using actin covalently labeled with N-(1-pyrenyl)iodoacetamide have been used to determine the effect of cytochalasin D on actin polymerization. A mechanism for the effect of cytochalasin D on actin polymerization is presented, which explains the experimental observation of a cytochalasin D-induced increase in the initial rate of polymerization and a decrease in the final extent of the reaction. Central to this mechanism is the Mg2+-dependent formation of cytochalasin D-induced dimers. The dimers serve as nuclei to enhance the polymerization rate. Binding of Mg2+ to a low affinity site on the dimer induces a conformational change which can be observed as a rapid fluorescence increase. A subsequent time-dependent fluorescence decrease observed prior to polymerization appears to represent ATP hydrolysis resulting in dissociation of the dimer and release of actin monomers containing ADP. We postulate that a slow rate of exchange of ATP for bound ADP relative to hydrolysis results in the accumulation of monomers containing ADP. As these monomers have a high critical concentration, the final extent of polymerization is reduced dramatically. The Mg2+ dependence of the final extent of polymerization in the presence of cytochalasin D is also explained in the context of this mechanism.  相似文献   

6.
G Panick  R Malessa  R Winter 《Biochemistry》1999,38(20):6512-6519
We examined the temperature- and pressure-induced unfolding and aggregation of beta-lactoglobulin (beta-Lg) and its genetic variants A and B up to temperatures of 90 degrees C in the pressure range from 1 bar to 10 kbar. To achieve information simultaneously on the secondary, tertiary, and quaternary structures, we have applied Synchrotron small-angle X-ray diffraction and Fourier transform infrared spectroscopy. Upon heating a beta-Lg solution at pH 7.0, the radius of gyration Rg first decreases, indicating a partial dissociation of the dimer into the monomers, the secondary structures remaining essentially unchanged. Above 50 degrees C, the infrared spectroscopy data reveal a decrease in intramolecular beta-sheet and alpha-helical structures, whereas the contribution of disordered structures increases. Within the temperature range from 50 to 60 degrees C, the appearance of the pair distance distribution function is not altered significantly, whereas the amount of defined secondary structures declines approximately by 10%. Above 60 degrees C the aggregation process of 1% beta-Lg solutions is clearly detectable by the increase in Rg and intermolecular beta-sheet content. The irreversible aggregation is due to intermolecular S-H/S-S interchange reactions and hydrophobic interactions. Upon pressurization at room temperature, the equilibrium between monomers and dimers is also shifted and dissociation of dimers is induced. At pressures of approximately 1300 bar, the amount of beta-sheet and alpha-helical structures decreases and the content of disordered structures increases, indicating the beginning unfolding of the protein which enables aggregation. Contrary to the thermal denaturation process, intermolecular beta-sheet formation is of less importance in pressure-induced protein aggregation and gelation. The spatial extent of the resulting protein clusters is time- and concentration-dependent. The aggregation of a 1% (w/w) solution of A, B, and the mixture AB results in the formation of at least octameric units as can be deduced from the radius of gyration of about 36 A. No differences in the pressure stability of the different genetic variants of beta-Lg are detectable in our FT-IR and SAXS experiments. Even application of higher pressures (up to 10 kbar) does not result in complete unfolding of all beta-Lg variants.  相似文献   

7.
D Salom  C Abad  L Braco 《Biochemistry》1992,31(34):8072-8079
We have investigated the conformational adaptability of gramicidin A incorporated into reverse micelles of sodium bis(2-ethylhexyl)sulfosuccinate (AOT)/isooctane/water, a so far unexplored "host" membrane-mimetic model system for this peptide. A high-performance liquid chromatographic strategy previously developed for the study of gramicidin in phospholipid vesicles and normal micelles [Ba?ó et al. (1989) FEBS Lett. 250, 67; Ba?ó et al. (1991) Biochemistry 30, 886] has been successfully extended to this system. The method has permitted the separation of peptide conformational species, namely, double-stranded dimers and monomers, and an accurate quantitation of their proportion in the inverted micellar environment. It has been demonstrated that, once inserted in the micelle, the double-stranded dimers undergo a dissociation process toward a thermodynamically stable monomeric configuration, whose monomerization rate constant (k1) is dependent in a bell-shaped manner on the water:surfactant mole ratio, w0. A tight correlation between k1 and the double-stranded dimer backbone conformation has been found from the comparison of chromatographic and circular dichroism data. In addition, fluorescence experiments indicate that the peptide tryptophans are in a rather nonpolar environment, with a restricted accessibility to water-soluble quenchers such as acrylamide.  相似文献   

8.
A comparative 270 MHz NMR spectroscopic study on the solution structure of the dimer d(TpT) 1, and its two analogues, namely, d(TpST) 2, and NH2d(TcmT) 4 has been reported. Analysis of chemical shifts and coupling constants indicate that: (i) The sugar moieties of the constituent nucleotides are not affected by modification of the internucleotide linkages and adopt preferentially an S-type conformation. (ii) The C4'-C5' bond in the pT part of the modified dimers 2 and 4 shows a large conformational freedom (gamma+ = 32% and 35%, respectively) compared to 1 (gamma+ = 75%). (iii) The population of the trans conformer about C5'-O5' is less important in d(TpST) 2 compared to d(TpT) 1. (iv) The C3'-O3' bond in 2 adopts a trans conformation as in 1. (v) The glycosidic bonds in the modified dimers 2 and 4 showed preferential syn conformation. UV and CD data show that the modified dimers 2 and 4 have poor tendency to stack intramolecularly, they also base pair less efficiently with d(ApA) as compared to d(TpT) 1.  相似文献   

9.
Conformational properties of (2'-5') and (3'-5') CpC have been determined by proton magnetic resonance spectroscopy at 220 MHz. The ribose ring structures are predominantly 3E with the exception of the ring from the 2'-phosphate fragment of C(2'-5')pC which exhibits an 2E pucker. Bases are oriented anti with respect to the ribose and the conformations about C4'-C5', C5'-O5', C3'-O3' (C2'-O2') are gg, g'g', and g+ in equilibrium g-, respectively. The dimers exist as mixtures of stacked (g+g+ and g-g- about the P-O(C) bonds) and unstacked species at 20 degrees C. Stacking is estimated to be 35% in both dimers.  相似文献   

10.
In living cells, P2Y(1) receptor dimerization was quantitated by an improved version of fluorescence resonance energy transfer donor photobleaching analysis. 44% of the P2Y(1) receptors expressed in HEK293 cell membranes exist as dimers in the resting state, inducible by agonist exposure to give 85-100% dimerization. Monomer and constitutive dimers are fully active. Agonist-induced dimerization follows desensitization and is fully reversible upon withdrawal of agonist. Receptor dimers are required for internalization at 37 degrees C but are not sufficient; at 20 degrees C dimerization also occurs, but endocytosis is abolished. Removal of the C-terminal 19 amino acids abolished both dimerization and internalization, whereas full activation by agonists was retained up to a loss of 39 amino acids, confirming active monomers. This receptor is known to bind through its last four amino acids (DTSL) to a scaffolding protein, Na/H exchanger regulatory factor-2, which was endogenous here, and DTSL removal blocked constitutive dimerization specifically. Distinction should therefore be made between the following: 1) constitutive dimers tethered to a scaffolding protein, together with effector proteins, within a signaling micro-domain, and 2) free dimers in the cell membrane, which here are inducible by agonist exposure. For the class A G-protein-coupled receptors, we suggest that the percentages of free monomers, and in many cases of induced free dimers, commonly become artifactually increased; this would arise from an excess there of the receptor over its specific scaffold and from a lack of the native targeting of the receptor to that site.  相似文献   

11.
The possible structural changes in the major isotype of parvalbumin from the toad (Bufo bufo japonicus) skeletal muscle caused by Ca2+ and Mg2+ binding have been analyzed by microcalorimetric titrations. Parvalbumin was titrated with Ca2+ in both the absence and presence of Mg2+ and with Mg2+ in the absence of Ca2+, at pH 7.0, and at 5 degrees, 15 degrees, and 25 degrees C. The two sites in a molecule were equivalent on Mg2(+)-Ca2+ exchange, but distinguishable on Ca2+ and Mg2+ binding. The reactions of parvalbumin with Ca2+ are exothermic at every temperature in both the absence and presence of Mg2+, but those with Mg2+ are always endothermic except for the binding to site 1 at 25 degrees C. The magnitudes of the hydrophobic and internal vibrational contributions to the heat capacity and entropy changes of parvalbumin on Ca2+ and Mg2+ binding and Mg2(+)-Ca2+ exchange have been estimated by the empirical method of Sturtevant [Sturtevant, J. M. (1977) Proc. Natl Acad. Sci. USA 74, 2236-2240]. Although no major conformational changes were noted between Ca2(+)- and Mg2(+)-bound forms of toad parvalbumin, the conformational difference was larger in Ca2+ (or Mg2+) binding to site 1 than site 2. This may indicate that the metal-free form is much less stable than any form with Ca2+ (or Mg2+) bound at one site at least. On Mg2(+)-Ca2+ exchange, the vibrational as well as hydrophobic entropy is only slightly increased in a parallel manner. In contrast, on Ca2+ (or Mg2+) binding, the hydrophobic entropy increases but the vibrational entropy decreases; the former indicates the sequestering of nonpolar groups from the surface to the interior of a molecule, and the latter suggests that the overall structures are tightened on Ca2+ (or Mg2+) binding but loosened on Mg2(+)-Ca2+ exchange. Despite the clear distinctions in the thermodynamic features, the conformational changes of toad parvalbumin are essentially the same as those of the two isotypes of bullfrog parvalbumins on Ca2+ binding and Mg2(+)-Ca2+ exchange.  相似文献   

12.
Structural alterations, as manifested by thermal transitions, caused by removal or binding of metal ions to human and bovine CuZn superoxide dismutases (SODs) were investigated by differential scanning calorimetry. Although holo forms of the two mammalian enzymes exhibited irreversible thermal transitions (delta Hcal. = 27.7 J/g and Td = 104 degrees C for bovine SOD; delta Hcal. = 23.6 J/g and Td = 101 degrees C for human SOD), only the bovine apoenzyme showed the presence of a less thermostable form (delta Hcal. = 10.7 J/g and Td = 63 degrees C). These observations suggested that human apo-SOD had considerably less conformational order than bovine apo-SOD. Reconstitution of human and bovine apoenzymes with Cu2+ and Zn2+ resulted in recovery of thermodynamic parameters and specific activity. Binding of Zn2+ alone to human apo-SOD resulted in the formation of two distinct structural units, detectable by differential scanning calorimetry, which underwent conformational disorder at 82 and 101 degrees C respectively. Saturation of binding sites with both Zn2+ and Cu2+ appeared to stabilize the enzyme structure further as shown by elimination of the low-temperature transition and the appearance of another thermal transition at a higher temperature.  相似文献   

13.
The rate of actin polymerization gradually decreased without changing the final level of polymerization, when incubated in the presence of 0.2 mM ATP at pH 8.0 and 25 degrees C. This change was much faster in Mg2+-actin than Ca2+-actin, and Mg2+-actin became denatured and unpolymerizable on prolonged incubation. The drop in the polymerization rate was due both to weakened nucleation and a slowed elongation rate in the incubated actin. The change in the polymerization rate was partially reversible by storing the sample at 0 degrees C. When the rate of polymerization dropped markedly on prolonged incubation, a gel filtration profile showed that Ca2+-actin existed as monomer not as oligomer. On the other hand, Mg2+-actin formed dimers, and other oligomers, as revealed by crosslinking analysis. There were changes in fluorescence intensities due to tyrosine and/or tryptophan residues of the actin molecule, and in difference absorption spectra, suggesting that conformational changes intermediate between native and denatured states occurred during incubation.  相似文献   

14.
Small-angle X-ray scattering data have been measured for rabbit skeletal muscle troponin C and its complexes with the venom peptides melittin and mastoparan as well as synthetic peptides based on regions of the troponin I sequence implicated in troponin C binding. At the neutral pH used in this study (pH 6.8), troponin C shows a tendency to form dimers in the presence of 4 mol equiv of Ca2+, but is monomeric in solution when 2 or less mol equiv of Ca2+ is present. The 4Ca2+.troponin C dimers dissociate upon binding melittin, mastoparan, and peptides based on residues 96-115, 1-30, and 1-40 in the troponin I sequence. This result suggests that the peptide-binding sites overlap with the regions of contact between troponin C molecules forming a dimer. Like the structurally homologous calcium-binding protein calmodulin, troponin C shows conformational flexibility upon binding different peptides. Upon binding melittin, troponin C contracts in a similar manner to calmodulin when it binds peptides known to form amphiphilic helices (e.g., melittin, mastoparan, or MLCK-I). In contrast, mastoparan binding to troponin C does not result in a contracted structure. The scattering data indicate troponin C also remains in an extended structure upon binding the inhibitory peptides having the same sequence as residues 96-115 in troponin I.  相似文献   

15.
16.
Thermodynamic parameters associated with the unfolding of the legume lectin, WBA II, were determined by isothermal denaturation. The analysis of isothermal denaturation data provided values for conformational stability and heat capacity for WBA II unfolding. To explore the role of intersubunit contact in stability, we carried out similar studies under identical conditions on Concanavalin A, a legume lectin of nearly similar size, buried hydrophobic surface area and tertiary structure to that of WBA II but with a different oligomerization pattern. Both proteins showed a reversible two-state unfolding with guanidine hydrochloride. As expected, the change in heat capacity upon unfolding was similar for both proteins at 3.5 and 3.7 kcal mol(-1) K(-1) for Concanavalin A and WBA II, respectively. Although the deltaG(H20) at the maximum stability of both proteins is around 16 kcal/mol, Concanavalin A exhibits greater stability at higher temperatures. The T(g) obtained for Concanavalin A and WBA II were 21 degrees C apart at 87.2 and 66.6 degrees C, respectively. The higher conformational stability at higher temperatures and the T(g) of Concanavalin A as compared to that of WBA II are largely due to substantial differences in the degree of subunit contact in these dimeric proteins. Ionic interactions and hydrogen bonding between the monomers of the two proteins also seem to play a significant role in the observed stability differences between these two proteins.  相似文献   

17.
A definition is proposed for biological ("internal") time tau(t) for a growing organism whose weight variation obeys the law w(t): tau(t) = 1/c(w) (t) = w(t)/w'(t), where t is physical ("external") time, w'(t) is weight increase rate, and c(w) (t) = w'(t)/w(t) is specific growth rate. Properties of functions tau(t) and w(tau) were studied for those cases when growth curves w(t) were described by Bertalanffy's or logistic equations.  相似文献   

18.
Deuterium NMR is used to study the structure and dynamics of the formyl C-2H bond in selectively deuterated gramicidin molecules. Specifically, the functionally different analogues 2HCO-Val1...gramicidin A and 2HCO-Val1-D-Leu2...gramicidin A are studied by 2H NMR so that any conformational or dynamical differences between the two analogues can be correlated with their difference in lifetime. These analogues are first synthesized, purified, and characterized and then incorporated into oriented bilayers of dimyristoylphosphatidylcholine sandwiched between glass coverslips. Phosphorous NMR line shapes obtained from these samples are consistent with the presence of the bilayer phase and indicate that the disorder exhibited by the lipid matrix is approximately of the same type and degree for both analogues. Deuterium NMR line shapes obtained from these samples indicate that the motional axis of the formyl group of gramicidin is parallel to the coverslip normal, that the distribution of motional axis orientations has a width of 7-9 degrees, and that a similar, major conformational and dynamical state exists for the formyl C-2H bond of both analogues. In this state, if the only motion present is fast axial rotation, then the experimentally derived angle between the formyl C-2H bond and the motional axis is consistent with the presence of a right-handed, single-stranded, beta 6.3 helical dimer but is not consistent with the presence of a left-handed, single-stranded, beta 6.3 helical dimer. However, if fast axial rotation is not the only motion present, then the left-handed, single-stranded, beta 6.3 helical dimer cannot be absolutely excluded as a possibility. Also, a second, minor conformational and dynamical state appears to be present in the spectrum of 2HCO-Val1-D-Leu2...gramicidin A but is not observed in the spectrum of 2HCO-Val1...gramicidin A. This minor conformational and dynamical state may reflect the presence of monomers, while the major conformational and dynamical state may reflect the presence of dimers.  相似文献   

19.
Quenched-flow mixing was used to characterize the kinetic behavior of the intermediate reactions of the skeletal muscle sarcoplasmic reticulum (SR) Ca-ATPase (SERCA1) at 2 and 21 degrees C. At 2 degrees C, phosphorylation of SR Ca-ATPase with 100 microM ATP labeled one-half of the catalytic sites with a biphasic time dependence [Mahaney, J. E., Froehlich, J. P., and Thomas, D. D. (1995) Biochemistry 34, 4864-4879]. Chasing the phosphoenzyme (EP) with 1.66 mM ADP 10 ms after the start of phosphorylation revealed mostly ADP-insensitive E2P (95% of EP(total)), consistent with its rapid formation from ADP-sensitive E1P. The consecutive relationship of the phosphorylated intermediates predicts a decrease in the proportion of E1P ([E1P]/[EP(total)]) with increasing phosphorylation time. Instead, after 10 ms the proportion of E1P increased and that of E2P decreased until they reached a constant 1:1 stoichiometry ([E1P]:[E2P] approximately 1). At 21 degrees C, phosphorylation displayed a transient overshoot associated with an inorganic phosphate (P(i)) burst, reflecting increased turnover of E2P at the higher temperature. The P(i) burst exceeded the decay of the EP overshoot, suggesting that rephosphorylation of the enzyme occurs before the recycling step (E2 --> E1). This behavior and the reversed order of accumulation of phosphorylated intermediates at 2 degrees C are not compatible with the conventional linear consecutive reaction mechanism: E1 + ATP --> E1.ATP --> E1P + ADP --> E2P --> E2.P(i) --> E1 + P(i). Solubilization of the Ca-ATPase into monomers using the nonionic detergent C(12)E(8) gave a pattern of phosphorylation in which E1P and E2P behave like consecutive intermediates. Kinetic modeling of the C(12)E(8)-solubilized SR Ca-ATPase showed that it behaves according to the conventional Ca-ATPase reaction mechanism, consistent with monomeric catalytic function. We conclude that the nonconforming features of native SERCA1 arise from oligomeric protein conformational interactions that constrain the subunits to a staggered or out-of-phase mode of operation.  相似文献   

20.
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