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.
2.
3.
4.
An engineered monomeric chorismate mutase (mMjCM) has been found to combine high catalytic activity with the characteristics of a molten globule. To gain insight into the dramatic structural changes that accompany binding of a transition-state analog, we examined mMjCM by isothermal calorimetry and compared it with its dimeric parent protein, MjCM (CM from Methanococcus jannaschii), a thermostable and conventionally folded enzyme. As expected for a ligand-induced ordering process, there is a large entropic penalty for binding to the monomer relative to the dimer (− TΔΔS = 5.1 ± 0.5 kcal/mol, at 20 °C). However, this unfavorable entropy term is largely offset by enthalpic gains (ΔΔH = − 3.5 ± 0.4 kcal/mol), presumably arising from tightening of non-covalent interactions throughout the monomeric complex. Stopped-flow kinetic measurements further reveal that the catalytic molten globule binds and releases ligands significantly faster than its natural counterpart, demonstrating that partial structural disorder can speed up molecular recognition. These results illustrate how structural plasticity may strongly perturb the thermodynamics and kinetics of transition-state recognition while negligibly affecting catalytic efficiency.  相似文献   

5.
6.
Rho termination factor is an essential hexameric helicase responsible for terminating 20-50% of all mRNA synthesis in Escherichia coli. We used single-molecule force spectroscopy to investigate Rho-RNA binding interactions at the Rho utilization site of the λtR1 terminator. Our results are consistent with Rho complexes adopting two states: one that binds 57 ± 2 nt of RNA across all six of the Rho primary binding sites, and another that binds 85 ± 2 nt at the six primary sites plus a single secondary site situated at the center of the hexamer. The single-molecule data serve to establish that Rho translocates 5′ → 3′ toward RNA polymerase (RNAP) by a tethered-tracking mechanism, looping out the intervening RNA between the Rho utilization site and RNAP. These findings lead to a general model for Rho binding and translocation and establish a novel experimental approach that should facilitate additional single-molecule studies of RNA-binding proteins.  相似文献   

7.
Despite the widespread presence of the globin fold in most living organisms, only eukaryotic globins have been employed as model proteins in folding/stability studies so far. This work introduces the first thermodynamic and kinetic characterization of a prokaryotic globin, that is, the apo form of the heme-binding domain of flavohemoglobin (apoHmpH) from Escherichia coli. This bacterial globin has a widely different sequence but nearly identical structure to its eukaryotic analogues. We show that apoHmpH is a well-folded monomeric protein with moderate stability at room temperature [apparent ΔG°UN(w) = − 3.1 ± 0.3 kcal mol− 1; mUN = − 1.7 kcal mol− 1 M− 1] and predominant α-helical structure. Remarkably, apoHmpH is the fastest-folding globin known to date, as it refolds about 4- to 16-fold more rapidly than its eukaryotic analogues (e.g., sperm whale apomyoglobin and soybean apoleghemoglobin), populating a compact kinetic intermediate (βI = 0.9 ± 0.2) with significant helical content. Additionally, the single Trp120 (located in the native H helix) becomes locked into a fully native-like environment within 6 ms, suggesting that this residue and its closest spatial neighbors complete their folding at ultrafast (submillisecond) speed. In summary, apoHmpH is a bacterial globin that shares the general folding scheme (i.e., a rapid burst phase followed by slower rate-determining phases) of its eukaryotic analogues but displays an overall faster folding and a kinetic intermediate with some fully native-like traits. This study supports the view that the general folding features of bacterial and eukaryotic globins are preserved through evolution while kinetic details differ.  相似文献   

8.
Escherichia coli HUαβ, a major nucleoid-associated protein, organizes chromosomal DNA and facilitates numerous DNA transactions. Using isothermal titration calorimetry, fluorescence resonance energy transfer and a series of DNA lengths (8 bp, 15 bp, 34 bp, 38 bp and 160 bp) we established that HUαβ interacts with duplex DNA using three different nonspecific binding modes. Both the HU to DNA molar ratio ([HU]/[DNA]) and DNA length dictate the dominant HU binding mode. On sufficiently long DNA (≥ 34 bp), at low [HU]/[DNA], HU populates a noncooperative 34 bp binding mode with a binding constant of 2.1 ± 0.4 × 106 M− 1, and a binding enthalpy of + 7.7 ± 0.6 kcal/mol at 15 °C and 0.15 M Na+. With increasing [HU]/[DNA], HU bound in the noncooperative 34 bp mode progressively converts to two cooperative (ω∼20) modes with site sizes of 10 bp and 6 bp. These latter modes exhibit smaller binding constants (1.1 ± 0.2 × 105 M− 1 for the 10 bp mode, 3.5 ± 1.4 × 104 M− 1 for the 6 bp mode) and binding enthalpies (4.2 ± 0.3 kcal/mol for the 10 bp mode, − 1.6 ± 0.3 kcal/mol for the 6 bp mode). As DNA length increases to 34 bp or more at low [HU]/[DNA], the small modes are replaced by the 34 bp binding mode. Fluorescence resonance energy transfer data demonstrate that the 34 bp mode bends DNA by 143 ± 6° whereas the 6 bp and 10 bp modes do not. The model proposed in this study provides a novel quantitative and comprehensive framework for reconciling previous structural and solution studies of HU, including single molecule (force extension measurement), fluorescence, and electrophoretic gel mobility-shift assays. In particular, it explains how HU condenses or extends DNA depending on the relative concentrations of HU and DNA.  相似文献   

9.
The ontogenetic changes of MAAs in the soft coral Heteroxenia fuscescens was studied in relation to their symbiotic state (azooxanthellate vs. zooxanthellate) under different temperature conditions in the Gulf of Eilat, northern Red Sea. The HPLC chromatograms for extracts of the planulae, azoo- and zooxanthellate primary polyps of H. fuscescens from all dates of collection yielded a single peak at 320 nm that has been identified as the compound palythine. Concentration of palythine in planulae at 23 °C was 7.57 ± 1 nmol mg− 1 protein and at 28 °C reached 17.29 ± 1 nmol × mg− 1 protein. Concentration of palythine in azooxanthellate primary polyps was 16.4 ± 3 nmol × mg− 1 protein and 28.37 ± 2.8 nmol × mg− 1 protein at 23 °C and 28 °C respectively. The palythine concentration for zooxanthellate primary polyps at 23 °C was 13 ± 3 nmol × mg− 1 protein and at 28 °C 32.7 ± 2 nmol mg− 1 protein. Palythine concentrations were significantly higher at 28 °C in the different animal groups and correlated linearly with the ambient collection temperature. This study shows for the first time that UVR and temperature act synergistically and affect the MAA levels of early life-history stages of soft corals.  相似文献   

10.
Ovigerous females of Cancer setosus are present year-round throughout most of its wide range along the Peruvian/Chilean Pacific coast (2°S-46°S). However, their number of egg-masses produced per year remains speculative and as such has neither been considered in latitudinal comparisons of reproduction, nor for its fisheries management. In order to reveal the effect of temperature on egg-mass production and egg-development, female C. setosus were held in through-flow aquaria under natural seasonal temperature conditions (16-23 °C) in Antofagasta (23°S), Northern Chile (05/2005-03/2006; 10 months), and at three constant temperatures (12, 16, 19 °C) in Puerto Montt (41°S), Central Southern Chile (09/2006-02/2007; 5 months). Female crabs uniformly produced up to 3 viable egg-masses within 4 1/2 months in Antofagasta and in Puerto Montt (at 19 °C). The second egg-mass was observed 62.5 days (± 7.6; N = 7) after the oviposition of the first clutch and a third egg-mass followed 73.5 days (± 12.5; N = 11) later in Antofagasta (at 16-23 °C). Comparably, a second oviposition took place 64.4 days (± 9.8, N = 5) after the first clutch and a third, 67.0 days (± 2.8, N = 2), thereafter, at 19 °C in Puerto Montt. At the two lower temperatures (16 and 12 °C) in Puerto Montt a second egg-mass was extruded after 82.8 days (± 28.9; N = 4) and 137 days (N = 1), respectively. The duration of egg-development from oviposition until larval hatching decreased from 65 days at 12.5 °C to 22.7 days at the observed upper temperature threshold of 22 °C. Based on the derived relationship between temperature and the duration of egg-development (y = 239.3175e− 0.107x; N = 21, r2 = 0.83) and data on monthly percentages of ovigerous females from field studies, the annual number of egg-masses of C. setosus was calculated. This analysis revealed an annual output of about one egg-mass close to the species northern and southern distributional limits in Casma (9°S) and Ancud (43°S), respectively, while at Coquimbo (29°S) about two and in Concepción (36°S) more than 3 egg-masses are produced per year.  相似文献   

11.
The factors controlling the stability, folding, and dynamics of integral membrane proteins are not fully understood. The high stability of the membrane protein bacteriorhodopsin (bR), an archetypal member of the rhodopsin photoreceptor family, has been ascribed to its covalently bound retinal cofactor. We investigate here the role of this cofactor in the thermodynamic stability and folding kinetics of bR. Multiple spectroscopic probes were used to determine the kinetics and energetics of protein folding in mixed lipid/detergent micelles in the presence and absence of retinal. The presence of retinal increases extrapolated values for the overall unfolding free energy from 6.3 ± 0.4 kcal mol− 1 to 23.4 ± 1.5 kcal mol− 1 at zero denaturant, suggesting that the cofactor contributes 17.1 kcal mol− 1 towards the overall stability of bR. In addition, the cooperativity of equilibrium unfolding curves is markedly reduced in the absence of retinal with overall m-values decreasing from 31.0 ± 2.0 kcal mol− 1 to 10.9 ± 1.0 kcal mol− 1, indicating that the folded state of the apoprotein is less compact than the equivalent for the holoprotein. This change in the denaturant response means that the difference in the unfolding free energy at a denaturant concentration midway between the two unfolding curves is only ca 3-6 kcal mol− 1. Kinetic data show that the decrease in stability upon removal of retinal is associated with an increase in the apparent intrinsic rate constant of unfolding, kuH2O, from ~1 × 10− 16 s− 1 to ~1 × 10− 4 s− 1 at 25 °C. This correlates with a decrease in the unfolding activation energy by 16.3 kcal mol− 1 in the apoprotein, extrapolated to zero SDS. These results suggest that changes in bR stability induced by retinal binding are mediated solely by changes in the activation barrier for unfolding. The results are consistent with a model in which bR is kinetically stabilized via a very slow rate of unfolding arising from protein-retinal interactions that increase the rigidity and compactness of the polypeptide chain.  相似文献   

12.
The occurrence of summer mortalities of the commercially important Pacific oyster, Crassostrea gigas, has increased in recent years. These mortality events occur during the late summer when water temperatures are at their highest. Many theories have been proposed concerning the causes including reproductive stress, environmental stress, disease, or synergistic interactions of these factors. C. gigas are grown intertidally and are exposed to the air (emersed) for hours at a time. These organisms can experience extreme changes in temperature during the course of a day. An oyster closed during emersion depletes the oxygen stores to near zero within the shell and builds up CO2 causing a decrease in tissue pH. The focus of this study is to determine the respiratory (pH, Po2, Pco2 and total CO2) and immune responses of oysters exposed to air at normal seasonal temperatures, and to determine whether these stresses associated with emersion inhibit the immune system of the oyster and contribute to the summer mortalities. The respiratory variables of the hemolymph of oysters submerged at 18 °C (pH = 7.52 ± 0.04 S.E.M., Po2 = 7.09 ± 0.53 S.E.M. kPa and Pco2 = 0.20 ± 0.03 S.E.M. kPa) varied significantly from oysters emersed for four hours at 22°C (pH = 7.11 ± 0.03 S.E.M., Po2 = 3.83 ± 0.15 S.E.M. kPa, Pco2 = 0.36 ± 0.03 S.E.M. kPa) and those emersed for four hours at 30 °C (pH = 6.84 ± 0.02 S.E.M., Po2 = 3.10 ± 0.12 S.E.M. kPa, Pco2 = 1.31 ± 0.06 S.E.M. kPa). The ability of hemocytes to kill the bacterium Vibrio campbellii was assessed using an in vitro assay to generate a killing index. There was no significant difference in the killing index between pH treatment groups (p = 0.856): at pH 7.6 killing index = 50.2% ± 2.33 S.E.M., at pH 6.6 killing index = 52.3% ± 3.67 S.E.M.. Temperature was the only factor to significantly affect the killing indices among temperature and oxygen treatment groups. The killing index was lowest (29.3% ± 3.25 S.E.M.) at 30 °C and 7% oxygen, simulating in vivo oxygen pressure in well-aerated conditions and 30 °C and 3% oxygen, simulating in vivo oxygen pressure in hypoxia (30.5% ± 3.25 S.E.M.), compared with the index in 7% oxygen at low temperature (18 °C) (44.4% ± 4.50 S.E.M.) or compared with low oxygen (3%) at low temperature (18 °C) (39.7% ± 2.51 S.E.M.). The seasonal and diurnal rise in temperature may, therefore, be an important factor contributing to summer mortalities of C. gigas.  相似文献   

13.
Sequence-resolved detection of pausing by single RNA polymerase molecules   总被引:4,自引:0,他引:4  
Herbert KM  La Porta A  Wong BJ  Mooney RA  Neuman KC  Landick R  Block SM 《Cell》2006,125(6):1083-1094
  相似文献   

14.
The aim of this study was to investigate if voluntary activation and force variability during maximal voluntary contraction (MVC) depends more on muscle (local) or body (core) temperature. Ten volunteers performed a 2-min MVC of the knee extensors under the control (CON) conditions (ambient temperature (21 °C), relative humidity (30%), and air velocity (∼0.1 m/s)) as well as after heating (HT) and cooling (CL) of the lower body. During water manipulation procedure lower body was immersed up to the waist in a water bath at ∼44 °C for 45 min for HT experiment, and ∼15 °C for 30 min for CL experiment. Peak torque, torque variability, muscle voluntary activation and half-relaxation time were assessed during the exercise. HT increased muscle (2.8±0.2 °C) and rectal (1.9±0.1 °C) temperatures while CL lowered muscle (2.2±0.2 °C) temperature, but did not affect rectal temperature. During 2-min MVC, peak torque decreased (P<0.05; SP>90%) and to a lower level in HT compared to CON and CL experiments (52.6±2.3% versus 69.0±2.3% and 65.6±1.9% MVC, respectively, P<0.05; SP>90%). Torque variability increased significantly during exercise and was significantly larger in HT and lower in CL compared to CON experiment. Voluntary activation of exercising muscle was more depressed in HT (i.e. greater central fatigue) and the smallest effect was found in CL compared to CON. In conclusion increased core and muscle temperature impairs voluntary activation and increases force variability of the exercising muscles while a local muscle cooling decrease force variability but has a small effect on central fatigue.  相似文献   

15.
16.
An early step in the morphogenesis of the double-stranded DNA (dsDNA) bacteriophage HK97 is the assembly of a precursor shell (prohead I) from 420 copies of a 384-residue subunit (gp5). Although formation of prohead I requires direct participation of gp5 residues 2-103 (Δ-domain), this domain is eliminated by viral protease prior to subsequent shell maturation and DNA packaging. The prohead I Δ-domain is thought to resemble a phage scaffolding protein, by virtue of its highly α-helical secondary structure and a tertiary fold that projects inward from the interior surface of the shell. Here, we employ factor analysis of temperature-dependent Raman spectra to characterize the thermostability of the Δ-domain secondary structure and to quantify the thermodynamic parameters of Δ-domain unfolding. The results are compared for the Δ-domain within the prohead I architecture (in situ) and for a recombinantly expressed 111-residue peptide (in vitro). We find that the α-helicity (∼ 70%), median melting temperature (Tm = 58 °C), enthalpy (ΔHm = 50 ± 5 kcal mol− 1), entropy (ΔSm = 150 ± 10 cal mol− 1 K− 1), and average cooperative melting unit (〈nc〉 ∼ 3.5) of the in situ Δ-domain are altered in vitro, indicating specific interdomain interactions within prohead I. Thus, the in vitro Δ-domain, despite an enhanced helical secondary structure (∼ 90% α-helix), exhibits diminished thermostability (Tm = 40 °C; ΔHm = 27 ± 2 kcal mol− 1; ΔSm = 86 ± 6 cal mol− 1 K− 1) and noncooperative unfolding (〈nc〉 ∼ 1) vis-à-vis the in situ Δ-domain. Temperature-dependent Raman markers of subunit side chains, particularly those of Phe and Trp residues, also confirm different local interactions for the in situ and in vitro Δ-domains. The present results clarify the key role of the gp5 Δ-domain in prohead I architecture by providing direct evidence of domain structure stabilization and interdomain interactions within the assembled shell.  相似文献   

17.
18.
We report on the response of reaction center (RC) from Rhodobacter sphaeroides (an archetype of membrane proteins) to the exposure at high temperature. The RCs have been solubilized in aqueous solution of the detergent N,N-dimethyldodecylamine-N-oxide (LDAO). Changes in the protein conformation have been probed by monitoring the variation in the absorbance of the bacteriochlorine cofactors and modification in the efficiency of energy transfer from tryptophans to cofactors and among the cofactors (through fluorescence measurements). The RC aggregation taking place at high temperature has been investigated by means of dynamic light scattering. Two experimental protocols have been used: (i) isothermal kinetics, in which the time evolution of RC after a sudden increase of the temperature is probed, and (ii) T-scans, in which the RCs are heated at constant rate. The analysis of the results coming from both the experiments indicates that the minimal kinetic scheme requires an equilibrium step and an irreversible process. The irreversible step is characterized by a activation energy of 205 ± 14 kJ/mol and is independent from the detergent concentration. Since the temperature dependence of the aggregation rate was found to obey to the same law, the aggregation process is unfolding-limited. On the other hand, the equilibrium process between the native and a partially unfolded conformations was found to be strongly dependent on the detergent concentration. Increasing the LDAO content from 0.025 to 0.5 wt.% decreases the melting temperature from 49 to 42 °C. This corresponds to a sizeable (22 kJ/mol at 25 °C) destabilization of the native conformation induced by the detergent. The nature of the aggregates formed by the denatured RCs depends on the temperature. For temperature below 60 °C compact aggregates are formed while at 60 °C the clusters are less dense with a scaling relation between mass and size close to that expected for diffusion-limited aggregation. The aggregate final sizes formed at different temperatures indicate the presence of an even number of proteins suggesting that these clusters are formed by aggregation of dimers.  相似文献   

19.
Cationic liposomes preferentially target tumor vasculature compared to vessels in normal tissues. The distribution of cationic liposomes along vascular networks is, however, patchy and heterogeneous. To target vessels more uniformly we combined the electrostatic properties of cationic liposomes with the strength of an external magnet. We report part I of development. We evaluated bilayer physical properties of our preparations. We investigated interaction of liposomes with target cells including the role of PEG (polyethylene-glycol), and determined whether magnetic cationic liposomes can respond to an external magnetic field. The inclusion of relatively high concentration of MAG-C (magnetite) at 2.5 mg/ml significantly increased the size of cationic liposomes from 105 ± 26.64 to 267 ± 27.43 nm and reduced the zeta potential from 64.55 ± 16.68 to 39.82 ± 5.26 mv. The phase transition temperature of cationic liposomes (49.97 ± 1.34 °C) reduced with inclusion of MAG-C (46.05 ± 0.21 °C). MAG-C cationic liposomes were internalized by melanoma (B16-F10 and HTB-72) and dermal endothelial (HMVEC-d) cells. PEG partially shielded cationic charge potential of MAG-C cationic liposomes, reduced their ability to interact with target cells in vitro, and uptake by major RES organs. Finally, application of external magnet enhanced tumor retention of magnetic cationic liposomes.  相似文献   

20.
Upstream interactions of Escherichia coli RNA polymerase (RNAP) in an open promoter complex (RPo) formed at the PR and PRM promoters of bacteriophage λ have been studied by atomic force microscopy. We demonstrate that the previously described 30-nm DNA compaction observed upon RPo formation at PR [Rivetti, C., Guthold, M. & Bustamante, C. (1999). Wrapping of DNA around the E. coli RNA polymerase open promoter complex. EMBO J., 18, 4464-4475.] is a consequence of the specific interaction of the RNAP with two AT-rich sequence determinants positioned from − 36 to − 59 and from − 80 to − 100. Likewise, RPos formed at PRM showed a specific contact between RNAP and the upstream DNA sequence. We further demonstrate that this interaction, which results in DNA wrapping against the polymerase surface, is mediated by the C-terminal domains of α-subunits (carboxy-terminal domain). Substitution of these AT-rich sequences with heterologous DNA reduces DNA wrapping but has only a small effect on the activity of the PR promoter. We find, however, that the frequency of DNA templates with both PR and PRM occupied by an RNAP significantly increases upon loss of DNA wrapping. These results suggest that α carboxy-terminal domain interactions with upstream DNA can also play a role in regulating the expression of closely spaced promoters. Finally, a model for a possible mechanism of promoter interference between PR and PRM is proposed.  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