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 93 毫秒
1.
A method is described for calculating r.b.e. values for normal tissues at risk in clinical neutron beam therapy. This is based on the assumption that with high l.e.t. radiations the slope of the cell survival curve is steeper, mainly in the initial or low-dose region. This effect is quantified by using two coefficients, one (epsilon) to produce a proportionate increase in the initial slope, and a second (eta) determining the change in the terminal slope (D0) of the survival curve. Analysis of published experimental data shows epsilon to be a variable quantity, different for different tissues; epsilon is larger when the survival curve has a large shoulder or slope ratio (rho). By contrast, eta is relatively constant (for a given beam) and less dependent on the tissue or end-point studied. For low doses, the r.b.e. approaches epsilon, which can be calculated given eta (characteristic of the beam) and rho (characteristic of the relevant tissue) [epsilon = eta + rho(eta - 1)]. This provides a useful approximation to the clinical r.b.e. for specific tissues relative to conventionally fractionated low-l.e.t. photons.  相似文献   

2.
Optimal monitor wavelengths and differential millimolar extinction coefficients (m delta epsilon) for rate determination of reactions catalyzed by adenosine deaminases on several substrates have been investigated as a function of pH in the range from 6.5 to 12. The values found are in some cases at variance with those quoted in the biochemical literature. The effect of pH on m delta epsilon values is shown to be clearly related to acid-base properties of product and/or substrate in the reaction. Experimental data are in most cases used to derive analytical functions describing the pH dependence of m delta epsilon. For the conversion of adenosine to inosine at pH 6.5, the following values of m delta epsilon +/- SE were obtained: at 263 nm, 8.27 +/- 0.02; at 264 nm, 8.36 +/- 0.02; at 265 nm, 8.27 +/- 0.03. These represent absolute maximal values as a function of pH.  相似文献   

3.
Adsorption mechanism of cadmium on juniper bark and wood   总被引:4,自引:0,他引:4  
In this study the capacity of sorbents prepared from juniper wood (JW) and bark (JB) to adsorb cadmium (Cd) from aqueous solutions at different pH values was compared. Adsorption behavior was characterized through adsorption kinetics, adsorption isotherms, and adsorption edge experiments. Results from kinetics and isotherm experiments showed that JB (76.3-91.6 micromol Cdg(-1) substrate) had 3-4 times higher adsorption capacity for Cd than JW (24.8-28.3 micromol Cdg(-1)). In addition to higher capacity, JB exhibited a higher strength of adsorption (45.3 versus 9.1 Lmmol(-1)) and faster uptake kinetics (0.0119 versus 0.0083 g micromol(-1)min(-1)) compared to JW. For both these adsorbents, increasing Cd adsorption with increasing solution pH in the range of 2-6 suggests that surface carboxyl groups (RCOOH) might be involved in interaction with Cd. Diffuse reflectance infrared Fourier transform (DRIFT) spectra showed that the surface concentration of carboxyl groups was higher on JB compared to JW. The ratio of Ca released to Cd adsorbed was 1.04 and 0.78 for JB and JW, respectively, indicating that Ca-Cd ion-exchange was the primary mechanism involved. The higher Ca content in JB (15 times more) and the surface RCOOH concentration (2.5 times more) can be attributed to the observed differences in Cd adsorption behavior between the two lignocellulosic adsorbents.  相似文献   

4.
The T-cell receptor (TCR) is a multisubunit complex consisting of the clonotypic Ti alpha and beta (or Ti gamma and delta) subunits and the invariant CD3 gamma, CD3 delta, CD3 epsilon, CD3 zeta, and CD3 eta subunits. Herein, we describe an additional product from the CD3 zeta/eta gene locus which we have termed CD3 theta. The cDNA derives from the first seven exons common to CD3 zeta and CD3 eta, 94 base pairs (bp) of the CD3 eta-specific exon 9 and an additional exon 10 encoding the carboxyl-terminal 15 amino acids and the 3'-untranslated region. The expression of CD3 theta is equivalent to that of CD3 eta in tissue distribution and level of expression as judged by RNase protection analysis. Despite the identity of the amino-terminal 121 amino acids of CD3 zeta, CD3 eta, and CD3 theta and an additional 31 amino acids shared between CD3 eta and CD3 theta, transfection of CD3 theta into the CD3 zeta- eta- T-cell hybridoma, MA5.8, failed to restore detectable surface TCR expression in contrast to transfection with CD3 zeta or CD3 eta. Analysis of the CD3 theta protein in transfectants indicated that CD3 theta is associated with the TCR intracellularly. However, unlike with CD3 zeta, Ti alpha-beta chains remain endoglycosidase H sensitive, suggesting a role for the unique COOH-terminal segment of CD3 theta in mediating TCR retention and/or degradation in a pre-Golgi compartment.  相似文献   

5.
The core of DNA polymerase III, the replicative polymerase in Escherichia coli, consists of three subunits (alpha, epsilon, and theta). The epsilon subunit is the 3'-5' proofreading exonuclease that associates with the polymerase (alpha) through its C-terminal region and theta through a 185-residue N-terminal domain (epsilon 186). A spectrophotometric assay for measurement of epsilon activity is described. Proteins epsilon and epsilon 186 and the epsilon 186.theta complex catalyzed the hydrolysis of the 5'-p-nitrophenyl ester of TMP (pNP-TMP) with similar values of k(cat) and K(M), confirming that the N-terminal domain of epsilon bears the exonuclease active site, and showing that association with theta has little direct effect on the chemistry occurring at the active site of epsilon. On the other hand, formation of the complex with theta stabilized epsilon 186 by approximately 14 degrees C against thermal inactivation. For epsilon 186, k(cat) = 293 min(-)(1) and K(M) = 1.08 mM at pH 8.00 and 25 degrees C, with a Mn(2+) concentration of 1 mM. Hydrolysis of pNP-TMP by epsilon 186 depended absolutely on divalent metal ions, and was inhibited by the product TMP. Dependencies on Mn(2+) and Mg(2+) concentrations were examined, giving a K(Mn) of 0.31 mM and a k(cat) of 334 min(-1) for Mn(2+) and a K(Mg) of 6.9 mM and a k(cat) of 19.9 min(-1) for Mg(2+). Inhibition by TMP was formally competitive [K(i) = 4.3 microM (with a Mn(2+) concentration of 1 mM)]. The pH dependence of pNP-TMP hydrolysis by epsilon 186, in the pH range of 6.5-9.0, was found to be simple. K(M) was essentially invariant between pH 6.5 and 8.5, while k(cat) depended on titration of a single group with a pK(a) of 7.7, approaching limiting values of 50 min(-1) at pH <6.5 and 400 min(-1) at pH >9.0. These data are used in conjunction with crystal structures of the complex of epsilon 186 with TMP and two Mn(II) ions bound at the active site to develop insights into the mechanisms of pNP-TMP hydrolysis by epsilon at high and low pH values.  相似文献   

6.
Qureshi SH  Moza B  Yadav S  Ahmad F 《Biochemistry》2003,42(6):1684-1695
The denaturation of bovine and horse cytochromes-c by weak salt denaturants (LiCl and CaCl(2)) was measured at 25 degrees C by observing changes in molar absorbance at 400 nm (Delta epsilon(400)) and circular dichroism (CD) at 222 and 409 nm. Measurements of Delta epsilon(400) and mean residue ellipticity at 409 nm ([theta](409)) gave a biphasic transition for both modes of denaturation of cytochromes-c. It has been observed that the first denaturation phase, N (native) conformation <--> X (intermediate) conformation and the second denaturation phase, X conformation <--> D (denatured) conformation are reversible. Conformational characterization of the X state by the far-UV CD, 8-anilino-1-naphthalene sulfonic acid (ANS) binding, and intrinsic viscosity measurements led us to conclude that the X state is a molten globule state. Analysis of denaturation transition curves for the stability of different states in terms of Gibbs energy change at pH 6.0 and 25 degrees C led us to conclude that the N state is more stable than the X state by 9.55 +/- 0.32 kcal mol(-1), whereas the X state is more stable than the D state by only 1.40 +/- 0.25 kcal mol(-1). We have also studied the effect of temperature on the equilibria, N conformation <--> X conformation and X conformation <--> D conformation in the presence of different denaturant concentrations using two different optical probes, namely, [theta](222) and Delta epsilon(400). These measurements yielded T(m), (midpoint of denaturation) and Delta H(m) (enthalpy change) at T(m) as a function of denaturant concentration. A plot of Delta H(m) versus corresponding T(m) was used to determine the constant-pressure heat capacity change, Delta C(p) (= ( partial differential Delta H(m)/ partial differential T(m))(p)). Values of Delta C(p) for N conformation <--> X conformation and X conformation <--> D conformation is 0.92 +/- 0.02 kcal mol(-1) K(-1) and 0.41 +/- 0.01 kcal mol(-1) K(-1), respectively. These measurements suggested that about 30% of the hydrophobic groups in the molten globule state are not accessible to the water.  相似文献   

7.
Tao Y  Zhang L  Yan F  Wu X 《Biomacromolecules》2007,8(7):2321-2328
Water-insoluble polysaccharide (TM3a), extracted from sclerotia of Pleurotus tuber-regium, was identified as a hyperbranched beta-d-glucan from the results of one- and two-dimensional NMR and GC-MS analysis. The degree of branching of TM3a is 65.5%. TM3a was fractionated by using a non-solvent addition method into 14 fractions, and its solution properties in 0.25 M LiCl/dimethylsulfoxide (DMSO) solution were studied systematically by using static laser light scattering, dynamic light scattering, and viscometry at 25 degrees C. The dependences among the values of intrinsic viscosity ([eta]), radius of gyration (z 1/2), and hydradynamic radius (Rh) on weight-average molecular weight (Mw) were found as the following: [eta] = 0.46Mw0.30+/-0.01, z 1/2 = 4.79 x 10-2Mw0.43+/-0.04, and Rh = 5.01 x 10-2Mw0.41+/-0.02 in the Mw range from 1.94 x 105 to 2.06 x 107 for TM3a in a 0.25 M LiCl/DMSO solution at 25 degrees C. The current theory of polymer solution was applied to explain the relationship among the fractal dimension, ratio of geometric to hydrodynamic radius (rho = z 1/2/Rh), and MwA2/[eta] of TM3a. The results indicated that TM3a existed as a compact chain conformation with a sphere-like structure in LiCl/DMSO solution. Furthermore, by using transmission electron microscopy, we observed directly the spherical molecules with an average diameter of 23.0 +/- 1.8 nm.  相似文献   

8.
The isoenzyme pattern of protein kinase C (PKC) in lymphocytes and airway smooth muscles (ASM) was examined by Western blot using commercially available monoclonal antibodies. The results showed the presence of PKC alpha, beta, gamma, epsilon, eta, mu and zeta in lymphocytes and PKC alpha, gamma, epsilon, eta and zeta in ASM. The unexpected feature was the presence of PKCgamma in both lymphocytes and ASM of guinea pigs. Expression of this PKC isoform is usually restricted to tissues in the central nervous system or spinal cord. Expression of PKC delta, theta, lambda and tau was not detected in either lymphocytes or ASM.  相似文献   

9.
The interactions between the N-terminal domain of the epsilon (epsilon186) and theta subunits of DNA polymerase III of Escherichia coli were investigated using electrospray ionization mass spectrometry. The epsilon186-theta complex was stable in 9 M ammonium actetate (pH 8), suggesting that hydrophobic interactions have a predominant contribution to the stability of the complex. Addition of primary alkanols to epsilon186-theta in 0.1 M ammonium acetate (pH 8), led to dissociation of the complex, as observed in the mass spectrometer. The concentrations of methanol, ethanol, and 1-propanol required to dissociate 50% of the complex were 8.9 M, 4.8 M, and 1.7 M, respectively. Closer scrutiny of the effect of alkanols on epsilon186, theta, and epsilon186-theta showed that epsilon186 formed soluble aggregates prior to precipitation, and that the association of epsilon186 with theta stabilized epsilon186. In-source collision-induced dissociation experiments and other results suggested that the epsilon186-theta complex dissociated in the mass spectrometer, and that the stability (with respect to dissociation) of the complex in vacuo was dependent on the solution from which it was sampled.  相似文献   

10.
Two conformationally restricted analogues of (-)-indolactam-V (1) (cis and trans amides) were examined for their binding selectivity to the synthetic C1 peptides of all protein kinase C (PKC) isozymes. Although the binding constants of the cis amide-restricted analogue (2) were equal to those of 1, the trans amide-restricted analogue (3) bound significantly only to the novel PKC (delta, epsilon, eta, theta) isozymes.  相似文献   

11.
The phylogenetic position of eleven 14-3-3 proteins from five protozoal species was tested relative to other eukaryotic 14-3-3 versions representing many of the previously described isoforms. The protozoal proteins, four from Entodinium caudatum, three from Entameoba histolytica and four from apicomplexan parasites formed clusters closer to the plant and animal epsilon isoforms than to the animal beta, gamma/eta, sigma/theta, and zeta isoforms. This extends the preliminary findings of Wang and Shakes (1996) but data from a wider range of genera are still required to strengthen our hypothesis that the protozoan isoforms may constitute novel isoforms of the 14-3-3 family.  相似文献   

12.
Sodium/Proton Exchange in Cultured Bovine Adrenal Medullary Cells   总被引:2,自引:1,他引:1  
We investigated the presence of Na+/H+ exchange in cultured bovine adrenal medullary cells. The intracellular pH in control cells measured by 5,5-dimethyl[2-14C]oxazolidine-2,4-dione was 7.13 +/- 0.02 (n = 6). Removal of Na+ from the incubation medium shifted the intracellular pH down to 6.67 +/- 0.12 (n = 6). Reintroduction of Na+ to the medium caused a rapid recovery in intracellular pH to 7.20-7.30 that was associated with an increase in uptake of 22Na+ by the cells. Both increases in intracellular pH and uptake of 22Na+ were inhibited by amiloride, an inhibitor of Na+/H+ exchange. The recovery of intracellular pH by addition of Na+ was partially inhibited by quinidine, another inhibitor of Na+/H+ exchange, but not by 4-acetamido-4'-isothiocyanostilbene-2,2'-disulfonic acid, an anion-exchange (Cl-/HCO3-) inhibitor. Li+ could substitute for Na+ in the recovery of intracellular pH. Carbachol caused an increase in intracellular pH from 7.12 +/- 0.01 to 7.21 +/- 0.02 (n = 10). This increase in intracellular pH caused by carbachol was inhibited by amiloride. These results suggest the existence of an amiloride-sensitive Na+/H+ exchange that regulates the intracellular pH in adrenal medullary cells.  相似文献   

13.
Continuous intracellular pH (pHi) measurements were performed in SIRC rabbit corneal epithelial cells using the pH-sensitive absorbance of intracellularly trapped 5(and 6)-carboxy-4',5'-dimethylfluorescein. Steady-state pHi in nominally bicarbonate free Ringer's solution averaged 6.87 +/- 0.02 (mean +/- S.E., n = 53). After intracellular acidification induced by the NH4Cl-prepulse technique, there was a sodium-dependent pHi recovery towards the normal steady-state pHi. The initial pHi recovery rate was a saturable function of extracellular sodium concentration with an apparent Km for external sodium of about 25 mM and a Vmax of about 0.28 pH units/min. Virtually no pHi recovery was observed in the absence of extracellular sodium. Sodium removal during steady state acidified the cells by 0.36 +/- 0.05 pH units (mean +/- S.E., n = 13) within 5 min. There was a dose-dependent inhibition of pHi recovery after NH4Cl prepulse by amiloride with an IC50 of about 15 microM. Amiloride in a concentration of 1 mM almost completely abolished pHi recovery. Amiloride (1 mM) applied during steady state induced an intracellular acidification of 0.2 +/- 0.03 pH units (mean +/- S.E., n = 7) within 5 min. These findings suggest that a Na+/H+ exchange is present in SIRC rabbit corneal epithelial cells. Na+/H+ exchange seems to be the major process involved in pHi recovery in SIRC cells after an intracellular acid load. Na+/H+ exchange also plays a role in the maintenance of steady-state pHi.  相似文献   

14.
D A Yphantis  T Arakawa 《Biochemistry》1987,26(17):5422-5427
Recombinant DNA derived human interferon gamma (IFN-gamma) from Escherichia coli was examined by equilibrium ultracentrifugation. Short-column equilibrium experiments at pH 6.9 in 0.1 M ammonium acetate buffer gave a z-average molecular weight of 33,500 +/- 1400 at infinite dilution, corresponding to 1.98 +/- 0.08 times the formula weight. Long- (2.6 mm) column experiments at pH 7.5 in 0.04 M imidazole buffer gave a molecular weight of 33,400 +/- 500. Under the latter conditions IFN-gamma behaves somewhat nonideally, with the departure from ideality accounted for by an effective (Donnan) charge of about 6+. No association of this dimer to form tetramer or higher polymers was observed, with the association constant for formation of tetramer from dimer K24 found to be less than 34 L mol-1. Similarly, no dissociation to monomers was observable, with the dissociation constant to monomer K21 being less than 5 X 10(-8) mol L-1. At pH 3.55 in 0.02 M buffer (acetate plus acetic acid), there was virtually complete dissociation of the dimer to monomer. Extreme nonideality was seen in this low ionic strength system, and the effective charge on the protein was estimated to be about 11+. The reduced molecular weight M(1 -upsilon rho) of the monomer was found to be about 4.09 +/- 0.20 kg mol-1; this corresponds to a molecular weight of 16,410 +/- 820, with the Scatchard definition of components. A small amount of a polymer with a molecular weight of about 0.5 X 10(6) was detected under these conditions.  相似文献   

15.
The rupture risk of abdominal aortic aneurysms (AAAs) is thought to be associated with increased levels of wall stress. Finite element analysis (FEA) allows the prediction of wall stresses in a patient-specific, non-invasive manner. We have recently shown that it is important to include the intra-luminal thrombus (ILT), present in approximately 70% of AAA, into FEA simulations of AAA. All FEA simulations to date assume an isotropic, homogeneous material behavior for this material. The purpose of this work was to investigate the multi-axial biomechanical behavior of ILT and to derive an appropriate constitutive relation. We performed planar biaxial testing on the luminal layer of nine ILT specimens obtained fresh in the operating room (9 patients, mean age 71+/-4.5 years, mean diameter 5.9+/-0.4 cm), and a constitutive relation was derived from this data. Peak stretch and maximum tangential modulus (MTM) values were recorded for the equibiaxial protocol in both the circumferential (theta) and longitudinal (L) directions. Stress contour plots were used to investigate the presence of mechanical anisotropy, after which an appropriate strain energy function was fit to each of the specimen datasets. The peak stretch values for the luminal layer of the ILT were (mean+/-SEM) 1.18+/-0.02 and 1.13+/-0.02 in the theta and L directions, respectively (p=0.14). The MTM values were 20+/-2 and 23+/-3N/cm(2) in the theta and L directions, respectively (p=0.37). From these results and our observation of the symmetry of the stress contour plots for each specimen, we concluded that the use of an isotropic strain energy function for ILT is appropriate. Each specimen data set was then fit to a second-order polynomial strain energy function of the first invariant of the left Cauchy-Green strain tensor, resulting in an accurate fit (average R(2)=0.92+/-0.02; range 0.80-0.99). Comparison of our previously reported, uniaxially derived constitutive relation with the biaxially derived relation derived here shows large differences in the predicted mechanical response, underscoring the importance of the appropriate experimental methods used to derive constitutive relations. Further work is merited in an effort to produce more accurate predictions of wall stresses in patient-specific AAA, and viscoelastic behaviors of the ILT.  相似文献   

16.
The flow properties of axoplasm have been studied in a defined chemical environment. Axoplasm extruded from squid giant axons was introduced into porous cellulose acetate tubes of diameter roughly equal to that of the original axon. Passage of axoplasm along the tube rapidly coated the tube walls with a layer of protein. By measuring the rate of low back and forth along the tube, the rheological properties of the axoplasm plug were investigated at a range of pressures and in a variety of media. Axoplasm behaves as a classical Bingham body the motion of which can be characterized by a yield stress (theta) and a plastic viscosity (eta p). In a potassium methanesulphonate medium containing 65 nM free Ca2+, theta averaged 109 +/- 46 dyn/cm2 and eta p1 146 +/- 83 P. These values were little affected by ATP, COLCHICINE, CYTOCHOLASIN B or by replacing K by Na but were sensitive to the anion composition of the medium. The effectiveness of different anions at reducing theta and eta p1 was in the order SCN greater than I greater then Br greater than Cl greater than methanesulphonate. Theta and eta p1 were also drastically reduced by increasing the ionized Ca. This effect required millimolar amounts of Ca, was unaffected by the presence of ATP and was irreversible. It could be blocked by the protease inhibitor TLCK. E.p.r. measurements showed that within the matrix of the axoplasm gel there is a watery space that is largely unaffected by anions or calcium.  相似文献   

17.
The regulation of intracellular pH (pHi) in rat sublingual mucous acini was monitored using dual-wavelength microfluorometry of the pH-sensitive dye BCECF (2',7'-biscarboxyethyl-5(6)-carboxyfluorescein). Acini attached to coverslips and continuously superfused with HCO3(-)-containing medium (25 mM NaHCO3/5% CO2; pH 7.4) have a steady-state pHi of 7.25 +/- 0.02. Acid loading of acinar cells using the NH4+/NH3 prepulse technique resulted in a Na(+)-dependent, MIBA-inhibitable (5-(N-methyl-N-isobutyl) amiloride, Ki approximately 0.42 microM) pHi recovery, the kinetics of which were not influenced by the absence of extracellular Cl-. The rate and magnitude of the pHi recovery were dependent on the extracellular Na+ concentration, indicating that Na+/H+ exchange plays a critical role in maintaining pHi above the pH predicted for electrochemical equilibrium. When the NH4+/NH3 concentration was varied, the rate of pHi recovery was enhanced as the extent of the intracellular acidification increased, demonstrating that the activity of the Na+/H+ exchanger is regulated by the concentration of intracellular protons. Switching BCECF-loaded acini to a Cl(-)-free medium did not significantly alter resting pHi, suggesting the absence of Cl-/HCO3- exchange activity. Muscarinic stimulation resulted in a rapid and sustained cytosolic acidification (t 1/2 < 30 sec; 0.16 +/- 0.02 pH unit), the magnitude of which was amplified greater than two-fold in the presence of MIBA (0.37 +/- 0.05 pH unit) or in the absence of extracellular Na+ (0.34 +/- 0.03 pH unit). The agonist-induced intracellular acidification was blunted in HCO3(-)-free media and was inhibited by DPC (diphenylamine-2-carboxylate), an anion channel blocker. In contrast, the acidification was not influenced by removal of extracellular Cl-. The Ca2+ ionophore, ionomycin, mimicked the effects of stimulation, whereas preloading acini with BAPTA (bis-(o-aminophenoxy)-ethane-N,N,N',N'-tetra-acetic acid) to chelate intracellular Ca2+ blocked the agonist-induced cytoplasmic acidification. The above results indicate that during muscarinic stimulation an intracellular acidification occurs which: (i) is partially buffered by increased Na+/H+ exchange activity; (ii) is most likely mediated by HCO3- efflux via an anion channel; and (iii) requires an increase in cytosolic free [Ca2+].  相似文献   

18.
Western blotting was used to identify the array of protein kinase C (PKC) isozymes expressed in the early (Day 4) and midcycle (Day 10) bovine corpus luteum (CL). PCKalpha, betaI, betaII, epsilon, and micro isozymes were detected in total protein samples prepared from both Day-4 and Day-10 corpora lutea. In contrast, specific antibodies for PKCgamma, eta, lambda, and theta isozymes failed to detect protein bands in the luteal samples. PKCbetaII and epsilon isozymes were expressed differentially at these two developmental stages of the bovine CL. In the Day-4 luteal samples, PKCepsilon was barely detectable; in contrast, in the Day-10 samples, the actin-corrected ratio for PKCepsilon was 1.16 +/- 0.13. This ratio was higher than the detected ratio for PKCbetaI and micro at this developmental phase of the CL (P < 0.01), but it was comparable with the ratio detected for the PCKalpha and betaII. The amount of PKCbetaII was, although not as dramatic, also greater in the Day-10 CL (actin-corrected ratio was 0.85 +/- 0.2) than in the Day-4 CL (0.35 +/- 0.09 [P < 0.01]). The actin-corrected ratios for all other PKC isozymes, alpha (Day 4 = 0.93 +/- 0.16, Day 10 = 0.97 +/- 0.09), betaI (Day 4 = 0.54 +/- 0.073, Day 10 = 0.48 +/- 0.74), and micro (Day 4 = 0.21 +/- 0.042, Day 10 = 0.21 +/- 0.38) were not different at these 2 days of the cycle. An experiment was designed to test whether activation of specific isozymes differed between CL that do or do not regress in response to PGF(2alpha). Bovine CL from Day 4 and Day 10 of the estrous cycle were collected and 1 mm CL fragments were treated in vitro for 0, 2.5, 5, 10 or 20 min with PGF(2alpha) (0.1, 1.0, and 10 nM) or minimal essential medium-Hepes vehicle. Translocation of PKC from cytoplasm to membrane fraction was used as indication of PKC activation by PGF(2alpha). Evidence for PKC activation was observed in both Day-4 and Day-10 luteal samples treated with 10 nM PGF(2alpha). Therefore, if PKC, an intracellular mediator associated with the luteal PGF(2alpha) receptor, contributes to the lesser sensitivity of the Day-4 CL, it is likely due to the differential expression of the epsilon and betaII isozymes of PKC at this stage and not due to an inability of the PGF(2alpha) receptor to activate the isozymes expressed in the early CL.  相似文献   

19.
In phylogenetic reconstructions by the parsimony method, utilizing 62 sequenced globin genes and pseudogenes (including 34 of the beta-globin gene family from eutherian orders Primates, Lagomorpha, Artiodactyla and Rodentia), the branch of primate psi beta pseudogenes and the goat embryonically expressed epsilon II gene group monophyletically together as orthologues of a common ancestral gene (labelled eta) distinct from orthologues of epsilon, gamma, delta and beta. This primate psi eta-goat eta branch is cladistically closer to epsilon and gamma than to delta and beta branches. In each eutherian order gene conversions replaced portions of delta by beta sequences, whereas in descent of Primates epsilon, gamma and eta mostly retained their separate ancient identities predating the radiation of Eutheria in all their exons and non-coding regions. The loci of the ancestral beta-globin gene cluster in basal eutherians and proto-primates, as deduced from beta-clusters representing the four eutherian orders, were linked 5'-epsilon-gamma-eta-delta-beta-3' with epsilon, gamma and eta being embryonically expressed genes, and delta and beta ontogenetically later expressed genes. Through deletions gamma was lost in artiodactyl evolution, eta in lagomorph and rodent evolution, and all DNA between exon 2 3' boundaries of eta and delta in prosimian lemuriform evolution (lemur having the hybrid pseudogene psi eta delta). Simian primates retained intact the five loci of the ancestral cluster. Not only did eta, after it became a pseudogene in the basal primates, persist intact in descent to present-day simians but in the line to hominoids it evolved during the last 40 million years at the decelerated rate of 1 X 10(-9) substitutions/site per year which is one-fifth the expected neutral rate. The possibility is suggested that the psi eta locus situated between fetal and adult chromosomal domains of the simian beta-globin gene cluster might play some role in a mechanism for ontogenetic switches of globin gene expression. However, not enough sequence data on genes and intergenic regions in DNA of species of primates and other mammals as yet exist to know if the slow rate of 1 X 10(-9) reflects the rate of a conserved functional gene or primarily reflects a decelerated neutral rate of hominoid DNA evolution, conceivably from enhanced DNA repair and longer generation times in hominoids. The further possibility is raised that gene correction (repair of damaged DNA that prevents emergence of new alleles) and gene conversion both more often involve strand copying of conserved than of rapidly evolving DNA.  相似文献   

20.
Enrichment and recovery of bovine serum albumin has been examined in a continuous foam separation column. The effects of the operating factors, superficial air velocity, feed flow rate, feed concentration and pH on the above characteristics was investigated. The protein enrichment decreased with the increase in the value of each of these parameters. Protein recovery increased with increasing air velocity, decreased with increasing feed flow rate and did not change very much with increasing feed concentration. Maximum protein recovery was obtained at the isoelectric point (pH 4.8) of the protein. Maximum protein recovery was found to be a strong function of the air velocity in the range 0.05-0.15 cm/s. Further increase in air velocity did not have much effect on recovery because of very large bubbles formed as a result of coalescence. Bubble size was determined as a function of the above factors in the liquid and foam sections of the column. It was found to be dependent on protein concentration, feed flow rate and solution pH. The effect was more significant in the foam section of the column. The bubbles in the foam section were significantly larger (about 3-10 times) than those in the liquid, with a sharp change at the foam-liquid interface. The bubble size measurements were used to calculate the interfacial area and it was shown that the rate of protein removal increases with increasing interfacial area.  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