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.
The dependence of initial rate v0 of ATP--PPi exchange reaction catalyzed by RNA-ligase of bacteriophage T4 on the concentration of ATP(s), pyrophosphate (z) and Mgcl2 has been determined. The dependence of v0 on s and z described by the equation v0 = k-1k2E0/(k-1 + K2) (1 + K1/s + k2/z) has been obtained for the reaction of E + S in equilibrium ES in equilibrium E1 + Z, where E--enzyme, E1--adenylylenzyme, S--ATP, Z--pyrophosphate, K1 and K2--constants of equilibrium, k-1, k2--velocity constants of transition of ES to E + S and E1 + Z, E0--complete concentration of enzyme. The low inhibition of the ATP--PPi exchange by the acceptor A(pA)2 and donors pAp, p(Ap)3, pCp has been shown. The dependence of v0 on the concentration of MgCl2 is consent with the incorporation of only dimagnesium salts of substrates in the isotope-exchange reaction.  相似文献   

2.
R Zhang  R Itri    M Caffrey 《Biophysical journal》1998,74(4):1924-1936
The variable-period x-ray standing wave (XSW) technique is emerging as a powerful tool for studying membrane structure. However, two significant problems arise when the method is used to characterize membranes of thickness dL < 100 A. First, the surface roughness, sigma(r), of the supporting reflecting mirror convolutes with the intrinsic half-width of the marker atom distribution in the membrane, sigma(in), and contributes to an apparent half-width, sigma, which is measured in the XSW experiment. Here we show how the latter terms are related quantitatively [sigma(in) = (sigma2 - sigma(r)2)(1/2)], such that rough mirrors give rise to larger marker atom distribution widths, sigma, and how the required quantity sigma(in) can be determined in the XSW measurement. Second, when the mean position of the marker atom layer, (z), is close to one or both boundaries of the membrane, its distribution function is truncated at the boundary. In such cases, we show why marker atom distribution should be expressed in terms of its first and second moments. We also demonstrate by numerical simulations of realistic samples how the physical parameters, sigma(r), sigma, (z), and dL, affect x-ray reflectivity and fluorescence yield profiles as an aid in their interpretation.  相似文献   

3.
AIM: To analyze whether bone mineral density (BMD) and bone resorption status are influenced by long-term metabolic control and duration of disease in adolescents with long-standing type 1 diabetes mellitus. METHODS: Twenty-seven adolescents (age 13.1 +/- 1.7 years, duration of diabetes 6.9 +/- 3.0 years) were studied. The BMD, expressed as z score, was measured at the lumbar spine (L1-L4) using dual-energy X-ray absorptiometry, while the urinary excretion of total deoxypiridinoline (Dpyd), a marker of bone resorption, was measured by immunoassay and was corrected by creatinine (Cr). Linear and multivariate correlations between lumbar BMD z score or Dpyd/Cr excretion and age and disease variables [short-term (Hb A(1c latest)) or long-term (Hb A(1c whole duration)) metabolic control, duration, 'diabetes impact index' (mean Hb A(1c whole duration) x duration of disease in months)] were sought. RESULTS: In diabetic subjects the mean BMD z score was -0.44 +/- (SD) 1.02 (95% CI: -0.03; -0.84), and the Dpyd/Cr excretion was not increased. A negative correlation was found between lumbar BMD z score and age (r -0.59; p = 0.001), duration (r -0.39; p = 0.04), and the diabetes impact index (r -0.4; p = 0.04). The Dpyd/Cr ratio correlated negatively with age (r -0.40; p = 0.04) and positively with height velocity (r 0.42; p = 0.04). By using multiple linear regression, age showed a significant inverse correlation with lumbar BMD z score (beta = -0.39; p = 0.0005). A negative correlation was found between lumbar BMD z score and Hb A(1c whole duration) (beta = -0.40; p = 0.02) or diabetes impact index (beta = -0.001; p = 0.01). CONCLUSIONS: Poor metabolic control may expose adolescents with long-standing type 1 diabetes to the risk of developing osteopenia in adult age. Optimization of metabolic control in growing diabetic children may prevent osteoporosis in later life.  相似文献   

4.
A formula by J. L. King gives the equilibrium mutation load as L = 2 sigma ui(1 - qi)/z - x) in which ui is the mutation rate to deleterious alleles at the ith locus, qi is the frequency of mutant alleles at this locus, x is the mean number of such mutant genes per individual before selection, z is the mean number in individuals eliminated by selection, and the summation is over all relevant loci. We show that this rule is inaccurate for intense selection and that a correct formula is L = 2 sigma ui(1 - qi) w/(z - x) = 2U w/(z - x) = 2U/(z - x + 2U) in which U is the mean number of new mutations per haploid genome in the population and w is the mean relative fitness before selection. If w/(z - x) less than 1/2, the mutation load is less than the Haldane value (U less than or equal to L less than or equal to 2U) and can be considerably less. In a diploid asexual population, however, with independent occurrence of mutations, L = 1 - e-2U regardless of the mode of selection.  相似文献   

5.
A basic set of equations describing the flows of volume (Jv) and solute (Js) across a leaky porous membrane, coupled to the differences of osmotic and hydrostatic pressures d pi and dP has been derived by using general frictional theory. Denoting the mean pore concentration of solute by c*s and the hydraulic and diffusive conductances by Lp and Ps/RT the equations take the form Jv = LpdP + sigma sLp d pi Js = c*s(1 - sigma f)Jv + Ps d pi/RT sigma s = theta (1 - DsVs/DwVw - Ds/Dos) sigma f = 1 - theta DsVs/DwVw - Ds/Dos in which Dw and Ds are the diffusion coefficients for water and solute in the pore and Dos that for free solution. The relation between the reflection coefficients sigma s and sigma f for osmosis and ultrafiltration is then given by sigma s = sigma f - (1- theta)(1 - Ds/Dos), where theta is the diffusive-driven:pressure-driven flow ratio. These equations follow from the fact that in leaky pores osmosis occurs by diffusion alone and that there cannot be any Onsager symmetry leading to sigma s = sigma f. Symmetry holds in the limits where either the pore is small, when sigma s = sigma f = 1, or where the pore is large when sigma s = sigma f = 0.  相似文献   

6.
The crystal structure of ethyl 2,3-dideoxy-alpha-D-erythro-hex-2-enopyranoside, C8H14O4, is orthorhombic, P2(1)2(1)2(1), with cell dimensions at 123 K [293 K] a = 11.220(2) [11.319(1)], b = 18.387(3) [18.458(2)], c = 8.509(2) [8.635(1)] A, Z = 8. There are two symmetry-independent molecules in the asymmetric unit. In both molecules, the conformation is oH5. The alkenic bond is almost exactly planar in one molecule, with C-1--C-2--C-3--C-4 = +0.8 degrees. In the other molecule, this torsion angle is +3.7 degrees. The glycosidic torsion angle, O-5--C-1--O-1--C-7, has normal exoanomeric values of +71 and +64 degrees. The conformation of the ethoxyl group is extended, with C-1--O-1--C-7--C-8 = +162 and +170 degrees. The primary alcohol group has different orientations, g/t on one molecule, g/g on the other. The characteristic glycosidic bond-shortening observed in the pyranosides is modified in this enopyranoside. Both the ring bond, O-5--C-1, and the glycosidic bond, C-1--O-1, are short, with distances ranging from 1.409 to 1.425 A. Solution and solid-state c.p.-m.a.s., 13C-n.m.r. spectra are reported.  相似文献   

7.
In contact with lipid bilayers and Ca2+-ions, the intracellular protein human annexin V (wild-type), Mr = 35,800, forms two types of cation-selective channels for the transport of Ca2+-, K+-, Na+- and Mg2+-ions, depending on the protein concentration [AN]. Type (I) channel events are large and predominant at high values [AN] > or = K = 5 nM at 296 K. At 50 mM Ca2+, symmetrical on both membrane sides, AN added at the cis side, the conductance is gCa(I) = 22 +/- 2 pS and at symmetrical 0.1 M K+-conditions: gK(I) = 32 +/- 3 pS, associated with two mean open-times tau1(I) = 0.68 +/- 0.2 ms and tau2(I) = 31 +/- 2 ms. Monoclonal anti-AN antibodies added to the trans-side first increase the mean open-times and then abolish the channel activity, suggesting that type (I) channels refer to a membrane spanning protein complex, probably a trimer T, which at [AN] > K changes its membrane organization to a higher oligomer, probably to the side-by-side double-trimer T2. The smaller type (II) channel events are predominant at low [AN] < or = K and refer to the (electroporative) adsorption complex of the monomer. The conductances g(i)(II) for symmetrical concentrations depend non-linearly on the voltage Um = Uext + U(AN), where U(AN) = 0.02 +/- 0.002 V is the electrostatic contribution of the Ca2+-AN complex and Uext the externally applied voltage. There is only one mean open-time tau(o)(II) which is voltage-dependent according to a functional of b x Um2 where b = 113.9 +/- 15 V(-2), yielding an activation Gibbs free energy of Ga = RT x b x Um2. The conformational flicker probability f(i)(II) in g(i)(II) = g(i)0(II) x gamma(i) x f(i)(II) is non-linearly voltage-dependent according to a functional of a x Um2. The Nernst term gamma(i) refers to asymmetrical ion concentrations. From a = 50 V(-2), independent of the ion type, we obtain f(i)0(II) = 0.03 +/- 0.002 and the conductances for the fully open-channel states: gCa0(II) = 69 +/- 3 pS (0.05 M Ca2+) and gK0(II) = 131 +/- 5 pS (1.2 M K+). From the electroporation term a = pi[r(p)2]epsilon0(epsilon(w) - epsilon(m))/(2 kTd) we determine the mean pore radius of the complex in its fully open state as r(p)= 0.86 +/- 0.05 nm. The adsorbed annexin V (Ca2+) monomer appears to electrostatically facilitate the electric pore formation at the contact interface between the protein and the lipid phase. The complex rapidly flickers and thus limits the ion transport in a voltage-dependent manner.  相似文献   

8.
Mechanical properties of trabecular bone. Dependency on strain rate.   总被引:8,自引:1,他引:7  
The effect of strain rate (epsilon) and apparent density (rho) on stiffness (E), strength (sigma u), and ultimate strain (epsilon u) was studied in 60 human trabecular bone specimens from the proximal tibia. Testing was performed by uniaxial compression to 5% specimen strain. Six different strain rates were used: 0.0001, 0.001, 0.01, 0.1, 1, and 10 s-1. Apparent density ranged between 0.23 and 0.59 g cm-3. Linear and non-linear regression analyses using strength, stiffness and ultimate strain as dependent variables (Y) and strain rate and apparent density as independent variables were performed using the following models: Y = a rho b epsilon c, Y = rho b(a + c epsilon; Y = (a + b rho)epsilon c, Y = a rho 2 epsilon c, E = a rho 3 epsilon c. The variations of strength and stiffness were explained equally well by the linear and the power function relationship to strain rate. The exponent was 0.07 in the power function relationship between strength and strain rate and 0.05 between stiffness and strain rate. The variation of ultimate strain was explained best using a power function relationship to strain rate (exponent = 0.03). The variation of strength and stiffness was explained equally well by the linear, power function and quadratic relationship to apparent density. The cubic relationship between stiffness and apparent density showed a less good fit. Ultimate strain varied independently of apparent density.  相似文献   

9.
T-cell receptor excision circles (TREC) may be a useful surrogate marker in HIV-1 infection for evaluating the likelihood of continued clinical stability and/or the response to therapeutics, including vaccines. Analysis of TREC in SHIV and SIV models of HIV-1 infection may provide additional information concerning the utility of TREC as a marker. We measured TREC in peripheral blood mononuclear cells (PBMC) from rhesus macaques in SHIV89.6p (n = 20) and SIVmac251 (n = 11) models of HIV-1 infection. TREC were also evaluated in tissues in the SIVmac251 model at end-point. In the SHIV89.6p model, TREC in PBMC were significantly lower at 12 weeks postinfection compared to preinfection levels. The decrease in TREC correlated with the decline in CD4+ T cells (r(s) = 0.496; P = 0.026), which in turn correlated inversely with serum viral loads at end-point (r(s) = -0.517; P = 0.019). Macaques that controlled SHIV89.6p infection to some degree (n = 6) had higher TREC at study end-point (P = 0.017). In the SIVmac251 model, TREC in PBMC were significantly reduced after 17 months of infection (P = 0.012) despite receiving highly active antiretroviral therapy (HAART) consisting of didanosine (ddI) and (R)-9-(2-phosphonylmethoxypropyl)-adenine (PMPA) when not cycling off therapy during scheduled treatment interruptions (STI). However, macaques that received continuous hydroxyurea (HU) in addition to the HAART regimen had higher end-point TREC compared to the non-HU group (P = 0.041), and the reduction in TREC observed at end-point within the HU group was not significant. In the SIVmac251 model, TREC correlated with the percentage of CD4+ T cells (r(s) = 0.426; P = 0.048) and CD4+CD28+ T cells (r(s) = 0.624; P = 0.002), and inversely with CD8+ T cells (r(s) = -0.622; P = 0.002), CD8+CD28- T cells (r(s) = -0.516; P = 0.014), and serum viral loads (r(s) = -0.627; P = 0.039). High levels of TREC were observed in the thymus, levels comparable to PBMC were seen in the lymph node, and low but detectable levels of TREC were present in bone marrow. The use of correlates of TREC as covariates in ANCOVA revealed that the decline in TREC in the SHIV 89.6p model reflected the decline in the percentage of CD4+ T-cells due to viral cytopathogenicity. In the SIVmac251 model, the decline in TREC was related to increased immune activation and proliferation due to viral replication, as reflected by decreases in percentages of CD4+CD28+ T cells and increases in CD8+ and CD8+CD28- T cells.  相似文献   

10.
We analyze the changes in the mean and variance components of a quantitative trait caused by changes in allele frequencies, concentrating on the effects of genetic drift. We use a general representation of epistasis and dominance that allows an arbitrary relation between genotype and phenotype for any number of diallelic loci. We assume initial and final Hardy-Weinberg and linkage equilibrium in our analyses of drift-induced changes. Random drift generates transient linkage disequilibria that cause correlations between allele frequency fluctuations at different loci. However, we show that these have negligible effects, at least for interactions among small numbers of loci. Our analyses are based on diffusion approximations that summarize the effects of drift in terms of F, the inbreeding coefficient, interpreted as the expected proportional decrease in heterozygosity at each locus. For haploids, the variance of the trait mean after a population bottleneck is var(delta(z)) = sigma(n)k=1 FkV(A(k)), where n is the number of loci contributing to the trait variance, V(A(1)) = V(A) is the additive genetic variance, and V(A(k)) is the kth-order additive epistatic variance. The expected additive genetic variance after the bottleneck, denoted (V*(A)), is closely related to var(delta(z)); (V*(A)) = (1 - F) sigma(n)k=1 kFk-1V(A(k)). Thus, epistasis inflates the expected additive variance above V(A)(1 - F), the expectation under additivity. For haploids (and diploids without dominance), the expected value of every variance component is inflated by the existence of higher order interactions (e.g., third-order epistasis inflates (V*(AA. This is not true in general with diploidy, because dominance alone can reduce (V*(A)) below V(A)(1 - F) (e.g., when dominant alleles are rare). Without dominance, diploidy produces simple expressions: var(delta(z)) = sigma(n)k=1 (2F)kV(A(k)) and (V(A)) = (1 - F) sigma(n)k=1 k(2F)k-1V(A(k)). With dominance (and even without epistasis), var(delta(z)) and (V*(A)) no longer depend solely on the variance components in the base population. For small F, the expected additive variance simplifies to (V*(A)) approximately equal to (1 - F)V(A) + 4FV(AA) + 2FV(D) + 2FC(AD), where C(AD) is a sum of two terms describing covariances between additive effects and dominance and additive X dominance interactions. Whether population bottlenecks lead to expected increases in additive variance depends primarily on the ratio of nonadditive to additive genetic variance in the base population, but dominance precludes simple predictions based solely on variance components. We illustrate these results using a model in which genotypic values are drawn at random, allowing extreme and erratic epistatic interactions. Although our analyses clarify the conditions under which drift is expected to increase V(A), we question the evolutionary importance of such increases.  相似文献   

11.
Batch metal sorption studies were conducted to compare the behavior of Gram-positive Bacillus subtilis and Gram-negative Escherichia coli as sorbents of Cd 2+ and Pb 2+ . A pH range from 3.0 to 6.5 was investigated at total metal concentrations of 1 2 10 -4.0 and 3.2 2 10 -5 M. Concentration apparent equilibrium sorption constants (K s n M ) and sorption capacity (S max n ) values were determined for the bacteria by fitting experimental data to one- ( n = 1) and two-site ( n = 2) Langmuir sorption isotherms. The sorption data for each of the bacteria were described well by a one-site model (r 2 > 0.9), Cd 2+ exhibited somewhat lower sorption affinities (log K s M =- 1.5 for B. subtilis , and -0.7 for E. coli ) than Pb 2+ (log K s M =-0.6 for B. subtilis and -0.8 for E. coli ). Corresponding S max values for Cd 2+ and Pb 2+ on B. subtilis were 0.36 mmole/g and 0.27 mmole/g, respectively. For E. coli Cd 2+ and Pb 2+ S max values were lower at 0.10 mmole/g and 0.21 mmole/g. A two-site sorption model yielded an improved fit for only the E. coli data with several orders of magnitude difference evident between high (Cd 2+ log K s1 M = 0.9; Pb 2+ log K s1 M = 1.5) and low (Cd 2+ log K s2 M =- 1.1; Pb 2+ log K s2 M = -1.6) affinity sorption sites. In addition, allowing for the presence of low affinity sorption (i.e., S max2 ) sites further increased the total E. coli metal sorption capacity closer to that of B. subtilis . As expected, the sorption of Cd 2+ and Pb 2+ by the bacteria exhibited a strong dependence on pH with sorption edges in the range of pH 4.2 to 5.6. The results of this study show that, despite differences in cell wall structure and composition, B. subtilis and E. coli exhibit remarkably similar sorption behavior toward Cd 2+ and Pb 2+ , respectively. These similarities can be attributed to the specific chemical reactivity of acidic functional groups (e.g., carboxyl, phosphoryl) that occur in the cell walls of both bacteria.  相似文献   

12.
Measurement of the osmotic response of a cell in the presence of cryoprotectant facilitates the determination of permeability coefficients which, in turn, can be used to design cryopreservation protocols which minimize osmotic stress. One problem encountered in determining permeability coefficients, using the Kedem-Katchalsky (K-K) model of membrane permeability, is that several combinations of the three passive coupled transport coefficients, namely, hydraulic permeability (L(p), microm min(-1) atm(-1)), solute permeability (P(gly), microm s(-1)), and the reflection coefficient (sigma), can give a similar fit to the measured data. A method for determining the "correct" set of coefficients is suggested. The osmotic response of 10 metaphase II mouse oocytes was measured on perfusion with 1.5 mol L(-1) glycerol at 24 degrees C. For 8 of 10 oocytes perfused, two combinations of L(p), P(gly), and sigma gave a predicted response which closely matched the measured osmotic response, depending upon the initial estimates supplied to the software for these parameters. For the remaining two oocytes, similar values for the permeability coefficients were generated regardless of the initial estimates. To determine the correct set of parameters, the K-K equations were used to predict experimental conditions for which volumetric histories would be distinctly different for the two sets of "best-fit parameters," and then additional experimental data were compared to these predictions. Thus a further three oocytes were perfused with 0.2 or 0.5 mol L(-1) glycerol in the absence of nonpermeating solute. In the presence of both 0.2 and 0.5 mol L(-1) glycerol, L(p) = 2.11 +/- 0.69, P(gly) = 0.0016 +/- 0.0015, and sigma = 0.44 +/- 0.11 yielded a very poor fit to the measured response while L(p) = 0.98 +/- 0.70, P(gly) = 0. 0031 +/- 0.0021, and sigma = 0.91 +/- 0.15 yielded a close fit to the measured response. Thus the latter combination of coefficients was taken to be correct.  相似文献   

13.
Myocardial Na+,K+-ATPase was studied in patients with aortic valve disease, and myocardial Na+,K+- and Ca2+-ATPase were assessed in spontaneously hypertensive rats (SHR) and hereditary cardiomyopathic hamsters using methods ensuring high enzyme recovery. Na+,K+-ATPase was quantified by [3H]ouabain binding to intact myocardial biopsies from patients with aortic valve disease. Aortic stenosis, regurgitation and a combination hereof were compared with normal human heart and were associated with reductions of left ventricular [3H]ouabain binding site concentration (pmol/g wet weight) of 56, 46 and 60%, respectively (p < 0.01). Na+,K+ and Ca2+-ATPases were quantified by K+- and Ca2+-dependent p-nitrophenyl phosphatase (pNPPase) activity determinations in crude myocardial homogenates from SHR and hereditary cardiomyopathic hamsters. When SHR were compared to age-matched Wistar Kyoto (WKY) rats an increase in heart-body weight ratio of 75% (p < 0.001) was associated with reductions of K+- and Ca2+-dependent pNPPase activities (mol/min/g wet weight) of 42 (p < 0.01) and 27% (p < 0.05), respectively. When hereditary cardiomyopathic hamsters were compared to age-matched Syrian hamsters an increase in heart-body weight ratio of 69% (p < 0.001) was found to be associated with reductions in K+- and Ca2+-dependent pNPPase activities of 50 (p < 0.001) and 26% (p = 0.05), respectively. The reductions in Na+,K+- and Ca2+-ATPases were selective in relation to overall protein content and were not merely the outcome of increased myocardial mass relative to Na+,K+- and Ca2+-pumps. In conclusion, myocardial hypertrophy is in patients associated with reduced Na+,K+-ATPase concentration and in rodents with reduced Na+,K+- and Ca2+-ATPase concentrations. This may be of importance for development of heart f in hypertrophic heart disease.  相似文献   

14.
Lee YS  Han JM  Kang T  Park YI  Kim HM  Kim S 《Molecules and cells》2006,21(2):213-217
Although AIMP1 (previously known as p43) is one of three auxiliary proteins bound to a macromolecular aminoacyl tRNA complex, it is also secreted as a cytokine controlling both angiogenesis and immune responses. Here we show that systemically administered purified recombinant human AIMP1 had anti-tumor activity in mouse xenograft models. In Meth A-bearing Balb/c mice, tumor volume increased about 28 fold in the vehicle treatment group, while an increase of about 16.7 fold was observed in the AIMP1-treated group. We also evaluated the anti-tumor activity of AIMP1 in combination with a sub-clinical dose of the cytotoxic anti-tumor drug, paclitaxel. The growth of NUGC-3 human stomach cancer cells was suppressed by 84% and 94% by the combinations of 5 mg/kg paclitaxel + 25 mg/kg AIMP1 (p = 0.03), and 5 mg/kg paclitaxel + 50 mg/kg AIMP1 (p = 0.02), respectively, while 5 mg/kg paclitaxel alone suppressed growth by only 54% (p = 0.02). A similar cooperative effect of AIMP1 and paclitaxel was observed in a lung cancer xenograft model. These results suggest that AIMP1 may be useful as a novel anti-tumor agent.  相似文献   

15.
The ligand binding problems on two-dimensional ladders, which model many important binding phenomena in molecular biology, are studied in details. The model is represented by four parameters, the interactions between ligands when bound to adjacent sites on opposite legs of the ladder (tau), the interactions between bound ligands in the longitudinal direction of the ladder (sigma), the number of binding sites that are covered by a bound ligand (m), and the intrinsic binding constant (K). The partition functions of ring ladders are approached with the transfer matrix method. A general relation is derived which connects the partition function of a linear ladder with that of a ring ladder. The results obtained apply to the general situation of multivalent binding, in which m>1. Special attention is paid to the case where the ligand covers one site (m=1). In this case explicit formulas are given for the partition functions of ring and linear ladders. Closed-form expressions are obtained for various properties of the system, including the degree of binding (theta), the midpoint in the binding isotherm (1/square root(tau sigma)), the initial and end slopes of the Scatchard plots (2sigma + tau - 4 and -sigma2 tau, respectively). From these closed-form formulas, sigma and tau may be extracted from experimental data. The model reveals certain features which do not exist in one-dimensional models. Using the general method discussed in [1], the recurrence relation is found for the partition functions. The analytical solution found for this model provides test cases to verify the numerical results for more complex two-dimensional models.  相似文献   

16.
Quantitative structure toxicity relationship (QSTR) equations were obtained to predict and describe the cytotoxicity of 31 phenols using logLD(50) as a concentration to induce 50% cytotoxicity of isolated rat hepatocytes in 2 h and logP as octanol/water partitioning: logLD(50) (microM)=-0.588(+/-0.059)logP+4.652(+/-0.153) (n=27, r(2)=0.801, s=0.261, P<1 x 10(-9)). Hydroquinone, catechol, 4-nitrophenol, and 2,4-dinitrophenol were outliers for this equation. When the ionization constant pK(a) was considered as a contributing factor a two-parameter QSTR equation was derived: logLD(50) (microM)=-0.595(+/-0.051)logP+0.197(+/-0.029)pK(a)+2.665(+/-0.281) (n=28, r(2)=0.859, s=0.218, P<1 x 10(-6)). Using sigma+, the Brown variation of the Hammet electronic constant, as a contributing parameter, the cytotoxicity of phenols towards hepatocytes were defined by logLD(50) (microM)=-0.594(+/-0.052)logP-0.552(+/-0.085)sigma+ +4.540(+/-0.132) (n=28, r(2)=0.853, s=0.223, P<1 x 10(-6)). Replacing sigma+ with the homolytic bond dissociation energy (BDE) for (X-PhOH+PhO.-->X-PhO.+PhOH) led to logLD(50) (microM)=-0.601(+/-0.066)logP-0.040(+/-0.018)BDE+4.611(+/-0.166) (n=23, r(2)=0.827, s=0.223, P<0.05). Hydroquinone, catechol and 2-nitrophenol were outliers for the above equations. Using redox potential and logP led to a new correlation: logLD(50) (microM)=-0.529(+/-0.135)logP+2.077(+/-0.892)E(p/2)+2.806(+/-0.592) (n=15, r(2)=0.561, s=0.383, P<0.05) with 4-nitrophenol as an outlier. Our findings indicate that phenols with higher lipophilicity, BDE, or sigma+ values or with lower pK(a) and redox potential were more toxic towards hepatocytes. We also showed that a collapse of hepatocyte mitochondrial membrane potential preceded the cytotoxicity of most phenols. Our study indicates that one or a combination of mechanisms; i.e. mitochondrial uncoupling, phenoxy radicals, or phenol metabolism to quinone methides and quinones, contribute to phenol cytotoxicity towards hepatocytes depending on the phenol chemical structure.  相似文献   

17.
18.
N-Isopropylammelide isopropylaminohydrolase, AtzC, the third enzyme in the atrazine degradation pathway in Pseudomonas sp. strain ADP, catalyzes the stoichiometric hydrolysis of N-isopropylammelide to cyanuric acid and isopropylamine. The atzC gene was cloned downstream of the tac promoter and expressed in Escherichia coli, where the expressed enzyme comprised 36% of the soluble protein. AtzC was purified to homogeneity by ammonium sulfate precipitation and phenyl column chromatography. It has a subunit size of 44,938 kDa and a holoenzyme molecular weight of 174,000. The K(m) and k(cat) values for AtzC with N-isopropylammelide were 406 micro M and 13.3 s(-1), respectively. AtzC hydrolyzed other N-substituted amino dihydroxy-s-triazines, and those with linear N-alkyl groups had higher k(cat) values than those with branched alkyl groups. Native AtzC contained 0.50 eq of Zn per subunit. The activity of metal-depleted AtzC was restored with Zn(II), Fe(II), Mn(II), Co(II), and Ni(II) salts. Cobalt-substituted AtzC had a visible absorbance band at 540 nm (Delta epsilon = 84 M(-1) cm(-1)) and exhibited an axial electron paramagnetic resonance (EPR) signal with the following effective values: g((x)) = 5.18, g((y)) = 3.93, and g((z)) = 2.24. Incubating cobalt-AtzC with the competitive inhibitor 5-azacytosine altered the effective EPR signal values to g((x)) = 5.11, g((y)) = 4.02, and g((z)) = 2.25 and increased the microwave power at half saturation at 10 K from 31 to 103 mW. Under the growth conditions examined, our data suggest that AtzC has a catalytically essential, five-coordinate Zn(II) metal center in the active site and specifically catalyzes the hydrolysis of intermediates generated during the metabolism of s-triazine herbicides.  相似文献   

19.
Tests and model selection for the general growth curve model   总被引:1,自引:0,他引:1  
J C Lee 《Biometrics》1991,47(1):147-159
The model considered here is a generalized multivariate analysis of variance model useful especially for many types of growth curve problems including biological growth and technology substitution. It is defined as Yp x N = Xp x m tau m x r Ar x N + epsilon p x N, where tau is unknown, and X and A are known design matrices of ranks m less than p and r less than N, respectively. Furthermore, the columns of epsilon are independent p-variate normal with mean vector 0 and common covariance matrix sigma. In general, p is the number of time (or spatial) points observed on each of the N cases, (m - 1) is the degree of polynomial in time, and r is the number of groups. The main focus of this paper is the selection of models for the general growth curve model with regard to the covariance matrix sigma. Likelihood ratio tests and selection procedures based on sample reuse and predictions are proposed. Special emphasis is on the serial covariance structure for sigma, which has been shown to be quite important in the prediction of biological data and technology substitution data. One-population and K-population problems are considered. Some of the results are illustrated with two sets of biological data.  相似文献   

20.
Two binuclear copper(II) complexes of 3,5-diisopropylsalicylic acid were characterized by single crystal X-ray diffraction methods and examined for anti-inflammatory activity using activated polymorphonuclear leukocytes and for anticonvulsant activities using electroshock and metrazol models of seizures. These complexes were crystallized from dimethylformamide (DMF) or diethylether. Tetrakis-mu-3,5-diisopropylsalicylatobis-dimethylformamidodicop per(II) [Cu(II)2(3,5-DIPS)4(DMF)2] I is in space group P 1; a = 10.393 (2), b = 11.258 (2), c = 12.734 (2) A, alpha = 96.64 (2), beta = 92.95 (2), gamma = 94.90 (2) degrees; V = 1471.7 (4) A3; Z = 1. Tetrakis-mu-3,5-diisopropylsalicylatobis-etheratodicopper(II ) [Cu(II)2(3,5-DIPS)4(ether)2] II is in space group P 1; a = 10.409 (3), b = 11.901 (4), c = 12.687 (6) A, alpha = 91.12 (5), beta = 90.84 (5), gamma = 100.90 (4) degrees; V = 1542 (1) A3; Z = 1. The structure of I was determined at 140 K from 4361 unique reflections (I > 2sigma(1)) and refined on F2 to R1 = 0.04 and wR2 = 0.09. The structure of II was determined at 180 K from 4605 unique reflections (I > 2sigma(I)) and refined on F2 to R1 = 0.05 and wR2 = 0.13. Each compound is a crystallographically centrosymmetric binuclear complex with Cu atoms bridged by four 3,5-diisopropylsalicylate ligands related by a symmetry center [Cu-Cu(i): 2.6139 (9) A in I and 2.613 (1) in II]. The four nearest O atoms around each Cu atom form a nearly rectangular planar arrangement with the square pyramidal coordination completed by the dimethylformamide (or diethylether) oxygen atom occupying an apical position, at a distance of 2.129 (2) A in I and 2.230 (3) A in II. Each Cu atom is displaced towards the DMF (or diethylether) ligand, by 0.189 A in I and 0.184 A in II, from the plane of the four O atoms. The crystal structures of I and II are essentially similar to each other, except for the DMF or diethylether accommodation. Many disorder phenomena were found in the crystal structure of I. Copper(II)2(3,5-DIPS)4(DMF)2 inhibited polymorphonuclear leukocyte (PMNL) oxidative metabolism in vitro. This effect was concentration related and significant for concentrations higher than 10 microg or 0.68 nmol/ml. Copper(II)2(3,5-DIPS)4(DMF)2 was more active than the parent ligand, 3,5-DIPS, as has been demonstrated with copper complexes of other non-steroidal anti-inflammatory drugs. The DMF and diethylether ternary complexes of Cu(II)2(3,5-DIPS)4 were found to have anticonvulsant activity in the maximal electroshock model of grand mal epilepsy in doses ranging from 26 to 258 micromol/kg of body mass following intraperitoneal, subcutaneous, or oral treatment. The DMF ternary complex was also found to be effective in the subcutaneous injection of metrazol model of petit mal epilepsy. We conclude that both ternary copper complexes are lipophilic and bioavailable, capable of facilitating the inflammatory response to brain injury and causing the subsidence of this response in bringing about remission of these disease states.  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