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 155 毫秒
1.
Phorbol ester treatment of granulocytes triggers release of superoxide (O) and a concomitant burst of DNA strand breaks. The relationship between the amount of O and the number of DNA breaks has not previously been explored. To quantify the relatively large amount of O generated over a 40-min period by 1 × 106 granulocytes/mL, a discontinuous “10-min pulse” method employing cytochrome c was used; 140 nmol O per 1 × 106 cells was detected. DNA strand breaks were quantified by fluorimetric analysis of DNA unwinding (FADU). To vary the level of O released by cells, inhibitors of the respiratory burst were used. Sodium fluoride (1–10 mM) and staurosporine (2–10 nM) both inhibited O production. In both cases, however, inhibition of strand breakage was considerably more pronounced than inhibition of O. Zinc chloride (50–200 μM) inhibited both O and DNA breaks, approximately equally. Dinophysistoxin-1 (okadaic acid) inhibited O production more effectively than it inhibited DNA breaks. O dismutes to H2O2, a reactive oxygen species known to cause DNA breaks. The addition of catalase to remove extracellular H2O2 had no effect on DNA breakage. Using pulse field gel electrophoresis, few double-stranded breaks were detected compared to the number detected by FADU, indicating that about 95% of breaks were single-stranded. The level of DNA breaks is not directly related to the amount of extracellular O or H2O2 in PMA-stimulated granulocytes. We conclude that either an intracellular pool of these reactive oxygen species is involved in breakage or that the metabolic inhibitors are affecting a novel strand break pathway. J. Cell. Biochem. 66:219–228, 1997. © 1997 Wiley-Liss Inc.  相似文献   

2.
The molecular theory of the previous paper in this series is extended to determine the effect of divalent metal ions on helix stability relative to coil at fixed ionic strength and nucleotide phosphate concentration. Specification of the state of condensed counterions, as well as their concentration, is essential for the solution of this problem, and it is assumed that they translate freely within a thin cylindrical shell close to the polynucleotide. As a function of divalent counterion concentration m the relative stability of the helix is highly nonlinear. Although the overall trend is that the helix stability increases with addition of divalent metal ion, there is a narrow concentration range for which it decreases slightly. The behavior of the relative stability as a function of m is determined by the translational degrees of freedom of the counterions, both univalent and divalent, both condensed and uncondensed. Detailed comparison of the theory with data is not given here, but it is pointed out that the calculated values of the relative stability are consistent with the order of magnitude of the observed effect Mg2+ on the melting temperature.  相似文献   

3.
A means to avoid the glucose effect in the production of baker's yeast from glucose and/or molasses in a fed batch culture by controlling the feed rate of fresh medium with an ad hoc measurement of the respiratory quotient, RQ, is presented. The feed rate is changed stepwise here such that the value of RQ ranges from 1.0 to 1.2 throughout the cultivation. Thus far, the specific growth rate based on the total cell mass and the growth yield obtained throughout are 0.24 hr?1 and 0.55 g cell/g glucose. Prior to the experimental run mentioned above, equations to predetermine the feed rate and concentration of glucose in the feed are derived from the mass balance of limiting substrates (glucose). Since values of either RQ or I (Q x, oxygen consumption rate with respect to the total cell mass in the fermenter) can be measured quite easily and reliably, computer control of the fermentation in light of this information is discussed.  相似文献   

4.
Trichoderma QM 9414 was aerobically grown on glucose as the sole carbon and energy sources in a chemostat culture. The specific rates of glucose consumption (QG), oxygen consumption (Q), and carbon dioxide production (Q) at the steady state were measured to estimate the growth and maintenance requirements. From the results it was estimated that 2 mol adenosine triphosphate (ATP) were produced when1 mol NADH was oxidized through the respiratory chain of this microorganism. The true growth yield for ATP (YATP) and specific ATP consumption rate for maintenance (Q) calculated with this value were 0.0106 g dry cell/mmol ATP and 5.2 mmol ATP/g dry cell/hr, respectively. Using the relationships between specific growth rate (μ) and (Q) and between μ and QG obtained from chemostat-culture data, cell and glucose concentration histories were estimated from the carbon dioxide production rate during the batch culture. The estimated cell concentrations agreed with the experimentally measured values. Glucose concentration were slightly overestimated.  相似文献   

5.
Alan E. Tonelli 《Biopolymers》1978,17(5):1175-1179
Solution conformations of the cyclic pentapeptide plant-hormone malformin A, whose conformational freedom is constrained by an intramolecular disulfide bridge, are derived and presented here. The nmr and CD data of Ptak are used to place restrictions on the search for possible malformin A solution conformers of low energy. Only two distinct conformers were found to be consistent with Ptak's data. Both structures are characterized by an internally buried (solvent-shielded) D -Cys2 amide proton, a seven-membered (1–3)hydrogen bond between (N–H) and (O?C), and a disulfide bridge conformation with a P chirality as manifested in the nmr study by the temperature independence of the amide proton chemical shifts for the D -Cys2 and D -Leu4 residues and the negative sign of the long wavelength maximum in the CD spectrum, respectively. Inspection of space-filling molecular models of both structures indicates severe steric barriers to their rapid interconversion. Thus, it appears that only one of the two conformers may be present in solution. The difference in their calculated dipole moments (4.6 and 6.9D) suggests an experimental method for distinguishing between the two proposed solution structures.  相似文献   

6.
The interaction between DNA and ionen polymers, -[N+(CH3)2(CH2)mN+(CH3)2(CH2)n], with m-n of 3–3, 6–6, and 6–10 were examined in order to know how the binding behavior of cationic polymers with DNA depends on the charge density of polycation. The ionen polymer has no bulky side chain and the binding forces with DNA would be attributed mainly to electrostatic interaction. When 3–3 ionen polymers were added to DNA solution, precipitable complexes with the ratio of cationic residue to DNA phosphate (+/?) of 1/1 and the free DNA molecules were segregated, while 6–6 and 6–10 ionen polymers formed soluble complexes with DNA molecules up to (+/?) = 0.5. This suggests that 3–3 ionen polymers bind cooperatively with DNA while 6–6 and 6–10 ionen polymers bind noncooperatively. The cooperative binding of 3–3 ionen polymer and the noncooperative binding of 6–6 ionen polymer were also supported by the thermal melting and recooling profiles from the midpoint between first and second meltings. It was concluded that the charge density of DNA phosphate is a critical value determining whether the ionen polymers bind to DNA by a cooperative or by a noncooperative binding, since the distance between successive cationic charges of 3–3 ionen polymer is shorter than that between successive phosphate charges on DNA double helix and those of 6–6 and 6–10 ionen polymers are longer.  相似文献   

7.
The use of methane for production of bacterial protein   总被引:3,自引:0,他引:3  
Single cell protein production was studied in a mixed bacterial culture grown in methane using batch and continuous culture techniques. Overall productivity was found to be higher in the continuous culture which gave a maximum productivity value p = 0.15 g/l/h. Methane and oxygen were consumed in the relation 1 : 1.7. Yield coefficients for methane, oxygen, and ammonium chloride were Y = 0.90, Y = 0.26, and Y = 0.14. The crude protein content of the biomass was 71%.  相似文献   

8.
Few environments challenge human populations more than high altitude, since the accompanying low oxygen pressures (hypoxia) are pervasive and impervious to cultural modification. Work capacity is an important factor in a population's ability to thrive in such an environment. The performance of work or exercise is a measure of the integrated functioning of the O2 transport system, with maximal O2 uptake (VO) a convenient index of that function. Hypoxia limits the ability to transport oxygen: maximal O2 uptake decreases with ascent to high altitude, and years of high altitude residence do not restore sea level VO values. Since Tibetans live and work at some of the highest altitudes in the world, their ability to exercise at very high altitude (<4,000 m) may define the limits of human adaptation to hypoxia. We transported 20 Tibetan lifelong residents of ≥4,400 m down to 3,658 m in order to compare them with 16 previously studied Tibetan residents of Lhasa (3,658 m). The two groups of Tibetans were matched for age, weight, and height. All studies were performed in Lhasa within 3 days of the 4,400 m Tibetans' arrival. Standard test protocol and criteria were used for attaining VO on a Monark bicycle ergometer, while measuring oxygen uptake (VO2, ml/kg − min STPD), heart rate (bpm), minute ventilation (VE, 1/min BTPS), and arterial oxygen saturation (Sa, %). The 4,400 m compared with 3,658 m residents had, at maximal effort, similar VO2 (48.5 ± 1.2 vs. 51.2 ± 1.4 ml/kg − min, P = NS), higher workload attained (211 ± 6 vs. 177 ± 7 watts, P < 0.01), lower heart rate (176 ± 2 vs. 191 ± 2 bpm, P < 0.01), lower ventilation (127 ± 5 vs. 149 ± 5 l/min BTPS, P < 0.01), and similar Sa(81.9 ± 1.0 vs. 83.7 ± 1.2%, P = NS). Furthermore, over the range of submaximal workloads, 4,400 m compared with 3,658 m Tibetans had lower VO2 (P < 0.01), lower heart rates (P < 0.01), and lower ventilation (P < 0.01) and Sa (P < 0.05). We conclude that Tibetans living at 4,400 m compared with those residing at 3,658 m achieve greater work performance for a given VO2 at submaximal and maximal workloads with less cardiorespiratory effort. Am J Phys Anthropol 105:21–31, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

9.
Quantification of blood fraction (fblood), blood oxygenation (S), melanin, lipofuscin and oxidised and reduced Cytochrome aa 3 and c was done from diffuse reflectance spectra captured in cortex, white matter, globus pallidus internus (GPi) and subthalamus during stereotactic implantations of 29 deep brain stimulation (DBS) electrodes with the aim of investigating whether the chromophores can give physiological information about the targets for DBS. Double‐sided Mann‐Whitney U ‐tests showed more lipofuscin in GPi compared to white matter and subthalamus (p < 0.05). Compared to the other structures, fbloodwas significantly higher in cortex (p < 0.05) and S lower in GPi (p < 0.05). Median values and range for fblood were 1.0 [0.2–6.0]% in the cortex, 0.3 [0.1–8.2]% in white matter, 0.2 [0.1–0.8]% in the GPi and 0.2 [0.1–11.7]% in the subthalamus. Corresponding values for S was 20 [0–81]% in the cortex, 29 [0–78]% in white matter, 0 [0–0]% in the GPi and 0 [0–92]% in the subthalamus. In conclusion, the measurements indicate very low oxygenation and blood volume for DBS patients, especially in the GPi. It would be of great interest to investigate whether this is due to the disease, the normal situation or an artefact of doing invasive measurements. (© 2013 WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim)  相似文献   

10.
A kinetic analysis of cytochrome P450-mediated desulfuration (activation) or dearylation (detoxication) showed that rat hepatic microsomes have a greater capacity to detoxify and a lower capacity to activate chlorpyrifos compared to parathion. Kinetic curves for the desulfuration of both parathion and chlorpyrifos were biphasic; K s of 0.23 and 71.3 μM were calculated for parathion, and 1.64 and 50.4 μM for chlorpyrifos. While phenobarbital (PB) exposure seemed to generally lower the Kmapp s for desulfuration except for the low Km activity on chlorpyrifos, the results were not statistically significant. While the low Km activity contributed 44 and 60% of the control Vmax for parathion and chlorpyrifos, respectively, it contributed 50 and 17% in PB-treated rats. These studies have indicated the presence of a low Km activity capable of functioning at very low substrate concentrations. A single dearylation K was calculated, 56.0 and 9.8 μM for parathion and chlorpyrifos, respectively. Phenobarbital exposure seemed to raise the Ks of dearylation; however, again, the results were not statistically significant. While numerous biochemical factors contribute to the overall toxicity levels of phosphorothionate insecticides, the in vitro efficiencies of hepatic microsomal desulfuration and dearylation of parathion and chlorpyrifos correspond to the acute toxicity levels.  相似文献   

11.
Assume k independent populations are given which are distributed according to R, …,Ri ∈ Θ ⊆ R ). Taking samples of size n the population with the smallest ϑ-value is to be selected. Using the framework of Le Cam's decision theory (Le Cam , 1986; Strasser , 1985) under mild regularity assumptions, an asymptotically optimal selection procedure is derived for the sequence of localized models. In the proportional hazards model with conditionally independent censoring, an asymptotically optimal adaptive selection procedure is constructed by substituting the unknown nuisance parameter by a kernel estimator.  相似文献   

12.
The 1H-nmr studies were extensively carried out to elucidate preferred conformations of dipeptides CH3C*O—X—NHCH3, with X = Abu, nVal, and Val in various solvents. The vicinal 1H—1H coupling constants for the NH—CαH moiety and those around the Cα—Cβ bond in the articulated side chain provided the information regarding the average conformation of these molecules. The results indicate that transformation of skeletal conformations takes place in solution among conformers having similar dihedral angles, θ, in the Karplus expression.  相似文献   

13.
A 20-membered cyclic peptide disulfide has been synthesized as a conformational model for disulfide loops of limited ring size. 1H-nmr studies at 270 MHz establish the presence of three intramolecular hydrogen bonds involving the Leu, Val, and methylamide NH groups in CDCl3. Evidence for peptide aggregation in CDCl3 is also presented. A structural transition involving loosening of the hydrogen bond formed by the Val NH group is observed upon the measured addition of (CD3)2SO to CDCl3. Hydrogen-bonding studies, together with unusually low field positions of the Cys(1) and Cys(6) CαH resonances and high J values provide support for an intramolecular antiparallel β-sheet conformation, facilitated by a chain reversal at the Aib-Ala segment. Extensive nuclear Overhauser effect studies provide compelling evidence for the proposed conformation and also establish a type I′ β-turn at the Aib-Ala residues, the site of the chain reversal.  相似文献   

14.
The 1H, 13C, and 15N resonances of FKBP when bound to the immunosuppressant, ascomycin, were assigned using a computer-aided analysis of heteronuclear double and triple resonance three-dimensional nmr spectra of [U-15N] FKBP/ascomycin and [U-15N, 13C] FKBP/ascomycin. In addition, from a preliminary analysis of two heteronuclear four-dimensional data sets, 3J coupling constants, amide exchange data, and the differences between the Cα and Cβ chemical shifts of FKBP to random coil values, the secondary structure of FKBP when bound to ascomycin was determined. The secondary structure of FKBP when bound to ascomycin in solution closely resembled the x-ray structure of the FKBP/FK506 complex but differed in some aspects from the structure of uncomplexed FKBP in solution. © 1993 John Wiley & Sons, Inc.  相似文献   

15.
Three freshwater and one marine algal species were grown under inorganic carbon limitation in laboratory continuous cultures. Comparisons were made between HCO3? alkalinity and bubbled CO2 as carbon sources. HCO3? alkalinity was an excellent source of inorganic carbon below specific pH levels, but chemical precipitation at high pH placed an upper limit on productivity that was far lower than potential light-limiting levels. With bubbled CO2 it was possible to achieve light limitation. The main factor controlling productivity was the mass flux of inorganic carbon added to the culture, which is the product of gas flow rate and influent P level. Small bubbles were more efficient than large bubbles at low gas flow rates and P levels, but led to froth flotation of algal cells and concomitant reductions in productivity at high bubble rates. At 1% CO2 productivity was still dependent on mass fluxes of added carbon, but was independent of bubble size. At high bubble rates with 1% CO2 narcosis was evident. Maximum yields occurred at intermediate dilution rates when inorganic carbon was supplied via bubbled gas.  相似文献   

16.
The 1H-nmr spectra of co-oligopeptides of tryptophan and glycine with structure H-Gly-Trp-(Gly)n-Trp-Gly-OH (n = 0–2) and those of several di- and tripeptides have been recorded at 360 MHz with CD3OD solutions containing 0.1N NaOD. The assignment of resonance signals was generally possible by comparing the spectra of structurally related peptides with each other. In order to solve the remaining ambiguities in the assignment, H-(αL,βS)(α,β-d2)Trp-OH, H-Trp-(αL,βS)(α,β-d2)Trp-OH, and H-Trp-(δ12232-d5)Trp-OH have been prepared and their spectra compared with those of the undeuterated compounds. The distribution of rotamers around the χ1 and (in two cases) χ2 torsion angles of the side chains has been obtained from the vicinal coupling constants 3J and from the long-range coupling constants 4J. These data and an analysis of the chemical shifts of the Gly-Cα protons suggest that the orientation of the aromatic side chain is influenced by the following order of decreasing interaction with the functional groups at N- and C-side: -NH2 > –NHCO– > –CONH–> –COO?. This rule does not hold for the second Trp residue of di- and tripeptides containing the -Trp-Trp- sequence, which has tentatively been attributed to steric effects.  相似文献   

17.
The relationship between cytosolic concentrations of Ca2+ (Ca) and Na+ (Na) were studied in preparations of rat submandibular and pancreatic acini loaded with the Ca2+-sensitive dye Fura-2 or the Na+-sensitive dye SBFI. Pancreatic acini showed no changes in Na during either transient or persistent changes in Ca. Increases in Ca produced by exposure of submandibular gland acini to carbachol, a muscarinic cholinergic agonist, were followed by an increase in Na after a delay of 5–10 s. When Ca2+ stores were mobilized without Ca2+ influx Na also increased, but in acini loaded with BAPTA, a nonfluorescent Ca2+ chelator, the transient increase in Ca2+ caused by mobilization of stored Ca2+ was virtually abolished, as was the increase in Na. In the presence of ionomycin, increases in Ca were followed by increases in Na. Ca2+-dependent increases in Na were abolished in Na+-free buffer and by the presence of furosemide, a blocker of Na+-K+-2Cl cotransport. In other studies, extracellular ATP (ATPo) produced an increase in Ca and Na. The steady-state increase in Ca was reduced by increasing extracellular Na+ concentrations (Na) in dose-dependent fashion (IC50 = 16.4 ± 4.7 mM Na+). Likewise, increasing Na reduced ATPo-stimulated 45Ca2+ uptake at steady state (IC50 = 15.8 ± 9.2 mM Na+). Changing Na had no effect on carbachol-stimulated increases in Ca. We conclude that, in rat submandibular gland acini, ATPo promotes an increase in Ca and Na via a common influx pathway and that, under physiologic conditions, Na+ significantly limits the ATPo-stimulated increase in Ca. In the presence of carbachol, however, Na rises in Ca-dependent fashion in submandibular gland acini via stimulation of Na+-K+-2Cl cotransport. © 1996 Wiley-Liss, Inc.  相似文献   

18.
DNA molecules condense into compact structures in the presence of a critical concentration of multivalent cations. To probe the contribution ofelectrostatic forces to condensation, we used mixtures of water with methanol (MeOH), ethanol (EtOH), and isopropanol (iPrOH) to vary the dielectric constant ? from 80 to 50. The condensation of pUC18 plasmids by hexaammine cobalt (III), Co(NH3), was monitored by total intensity and dynamic light scattering, electron microscopy, andCD. The total scattering intensity increased as ? went from 80 to 70, and then decreased as ? decreased further. Ultraviolet spectrophotometry confirmed that the loss of intensity at low ? was not due to the particles' settling out of solution. The rate as well as the extent of condensation increased as? was lowered from 80 to 70, and also depended on the species of alcohol (MeOH < EtOH < iPrOH). The hydrodynamic radii RH of the particles, however, remained roughly the same at 300–350 A and was independent of the species of alcohol. RH increased below ? = 70. The critical concentration of Co(NH3) required to induce DNA condensation decreased from 21 μM to about 16 μM as the dielectric constant decreased from 80 to 70, and decreased moderately with the nonpolarity of the alcohol. The fraction of DNA charge neutralized at the onset of DNA condensation was calculated by a modification of Manning's two-variable counterion condensation theory to be 0.90 ± 0.01, independent of ?. By electron microscopy we observed that the condensed particles changed from about 93% toroids at ? = 80 to 89% rods at ? = 70 and 98% rods at ? = 65. At epsi; lower than 65, DNA collapsed into a network of multistranded fibers. The morphology of condensed DNA particles, whether toroids, rods, or fibers, was independent of the alcohol species. CD spectra in ethanol–water mixtures indicated that both closed circular and linearized plasmids were in the B conformation when condensed with Co(NH3)3+6 at ?≥ 70, although the closed circular molecules exhibited a weak Ψ-DNA spectrum. A transition from the B to A formtook place between ? = 70 and 60, well above the normal dielectric constant of ? = 40 for this transition, indicating that ethanol and Co(NH3) synergistically promote the B–A transition. We interpret these results to mean that alcohols have both electrostatic and structural effects on DNA, leading to three regimes of condensation. At the lowest alcohol concentrations the B conformation is stableand condensation is relatively slow, allowing time for the packing adjustments necessary to form toroids. At intermediate alcohol concentrations condensation is faster, and the combined effects of solvent and Co(NH3) locally destabilize the double helix, permitting DNA foldbacks that lead to rodlike condensates. Rods become shorter as wellas more numerous as ? decreases from 80 to 65–60, indicatingincreasing destabilization as alcohol increases. At the lowest dielectric constants, alcohol and Co(NH3) produce A-DNA, which strongly self-adheres and rapidly aggregates intofibrous networks, not allowing time for more compact condensates to form. © 1995 John Wiley & Sons, Inc.  相似文献   

19.
A survey has been made, by using computer methods, of the types of helices which polypeptide chains can form, taking into account steric requirements and intramolecular hydrogen-bonding interactions. The influence on these two requirements, of small variations in the bond angles of the peptide residues, or of small changes in the overall dimensions of the helix (pitch and residues per turn), have been assessed for the special case of the α-helix. Criteria for the formation of acceptable hydrogen bonds have also been applied to helices of other types, viz., the 3, γ?, ω?, and π-helices. It was shown that the N? H … O and H … O? C angles in hydrogen bonds are sensitive to changes in either the NCαC′ bond angle or in the rotational angles about the N? Cα and Cα? C′ bonds. However, the variants of the α-helix observed experimentally in myoglobin can all be constructed without distortion of the hydrogen bonds. For α-helices, the steric and hydrogen bonding requirements are more easily fulfilled with an NCαC′ bond angle of 111°, rather than 109.5°. The decreased stability observed for the left-handed α-helix relative to the right-handed one for L -amino acids is due essentially only to interactions of the Cβ atom of the side chains with atoms in adjacent peptide units in the backbone, and interactions with atoms in adjacent turns of the helical backbone are not significantly different in the two helices. Restrictions in the freedom of rotation of bulky side chains may have significant kinetic effects during the formation of the α-helix from the “random coil” state.  相似文献   

20.
The kinetics of ethidium's intercalative binding to DNA packaged in bacteriophage T7 and two T7 deletion mutants have been determined, using enhancement of fluorescence to quantitate binding. At a constant ethidium concentration, the results can be described as first-order binding with two different rate constants, k (= k1 + k?1) and k (= k2 + k?2). The larger rate constant (k) was at least four orders of magnitude smaller than the comparable first-order forward rate constant for binding to DNA released from its capsid. At 25°C values of k decreased as the amount of DNA packaged per internal volume increased. This latter observation indicates that the rate of ethidium's binding to packaged T7 DNA is limited by an event that occurs inside of the DNA-containing region of T7, not by the crossing of T7 capsid's outer shell. Arrhenius plots of kM are biphasic, indicating a transition for packaged DNA at a temperature of 20°C. The data indicate that k s are limited by either sieving of ethidium during its passage through the packaged DNA or subsequent hindered intercalation.  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