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.
In spinal deformation studies, three-dimensional reconstruction of the spine is frequently represented as a curve in space fitted to the vertebral centroids. Conventional interpolation techniques such as splines. Bezier and the least squares method are limited since they cannot describe precisely the great variety of spinal morphologies. This article presents a more general technique called dual kriging, which includes two mathematical constituents (drift and covariance) to adjust the interpolated functions to spinal deformity better. The cross-validation technique was used to compare the parametric representations of spinal curves with different combinations of drift and covariance functions. Model validation was performed from a series of analytic curves reflecting typical scoliotic spines. Calculation of geometric torsion, a sensitive parameter, was done to evaluate the accuracy of the kriging models. The best model showed an absolute mean difference of 1.2 x 10~5 (±7·1 × l()~ 5) mm?1 between the analytical and estimated geometric torsions compared to 5·25 × 10~ (±3.7 × 310~2) mm* 1 for the commonly used least-squares Fourier series method, a significant improvement in spinal torsion evaluation.  相似文献   

2.
Novel magnetic resonance imaging sequences have and still continue to play an increasing role in neuroimaging and neuroscience. Among these techniques, diffusion-weighted imaging (DWI) has revolutionized the diagnosis and management of diseases such as stroke, neoplastic disease and inflammation. However, the effects of aging on diffusion are yet to be determined. To establish reference values for future experimental mouse studies we tested the hypothesis that absolute apparent diffusion coefficients (ADC) of the normal brain change with age. A total of 41 healthy mice were examined by T2-weighted imaging and DWI. For each animal ADC frequency histograms (i) of the whole brain were calculated on a voxel-by-voxel basis and region-of-interest (ROI) measurements (ii) performed and related to the animals' age. The mean entire brain ADC of mice <3 months was 0.715(+/-0.016) x 10(-3) mm2/s, no significant difference to mice aged 4 to 5 months (0.736(+/-0.040) x 10(-3) mm2/s) or animals older than 9 months 0.736(+/-0.020) x 10(-3) mm2/s. Mean whole brain ADCs showed a trend towards lower values with aging but both methods (i + ii) did not reveal a significant correlation with age. ROI measurements in predefined areas: 0.723(+/-0.057) x 10(-3) mm2/s in the parietal lobe, 0.659(+/-0.037) x 10(-3) mm2/s in the striatum and 0.679(+/-0.056) x 10(-3) mm2/s in the temporal lobe. With advancing age, we observed minimal diffusion changes in the whole mouse brain as well as in three ROIs by determination of ADCs. According to our data ADCs remain nearly constant during the aging process of the brain with a small but statistically non-significant trend towards a decreased diffusion in older animals.  相似文献   

3.
Biogenic polyamines, putrescine, spermidine, and spermine, are ubiquitous cellular cations and exert multiple biological functions. Polyamine analogues mimic biogenic polyamines at macromolecular level but are unable to substitute for natural polyamines and maintain cell proliferation, indicating biomedical applications. The mechanistic differences in DNA binding mode between natural and synthetic polyamines have not been explored. The aim of this study was to examine the interaction of calf thymus DNA with three polyamine analogues, 1,11-diamino-4,8-diazaundecane (333), 3,7,11,15-tetrazaheptadecane x 4 HCl (BE-333), and 3,7,11,15,19-pentazahenicosane x 5 HCl (BE-3333), using FTIR, UV-visible, and CD spectroscopy. Polyamine analogues bind with guanine and backbone PO2 group as major targets in DNA, whereas biogenic polyamines bind to major and minor grooves as well as to phosphate groups. Weaker interaction with DNA was observed for analogues with respect to biogenic polyamines, with K(333) = 1.90 (+/-0.5) x 10(4) M(-1), K(BE-333) = 6.4 (+/-1.7) x 10(4) M(-1), K(BE-3333) = 4.7 (+/-1.4) x 10(4) M(-1) compared to K(Spm) = 2.3 (+/-1.1) x 10(5) M(-1), K(Spd) = 1.4 (+/-0.6) x 10(5) M(-1), and K(Put) = 1.02 (+/-0.5) x 10(5) M(-1). A partial B- to A-DNA transition was also provoked by analogues. These data suggest distinct differences in the binding of natural and synthetic polyamines with DNA.  相似文献   

4.
Short peptides sequences were selected that showed binding selectivity towards healthy or oxidised (unhealthy) low density lipoprotein (LDL), respectively. These were investigated for application in atherosclerosis risk monitoring. Comparison was also made with the LDL receptor ligand repeat peptide (LR5). The peptides were immobilised on a gold surface plasmon resonance surface and LDL binding detected as a shift in the resonance. 3.7x10(7) (+/-5.6x10(6)) LDL/mm(2)/microg/ml solution LDL were bound on GlySerAspGlu-OH and 6.8x10(7) (+/-9.2x10(6)) LDL/mm(2)/microg/ml on GlyCystineSerAspGlu, compared with approximately 10(8) LDL/mm(2)/microg/ml on LR5. In this first group, binding of LDL decreased with oxidation level and a good correlation was found between LDL binding and residual amino groups on the apoprotein of the LDL following oxidation, or the change in relative electrophoretic mobility (REM) of LDL. The decrease in binding was 1.1x10(7) LDL particles/mm(2) per% oxidation for GlySerAspGlu-OH, 1.8x10(7) LDL particles/mm(2) per% oxidation for GlyCystineSerAspGlu and 2.4x10(7) LDL particles/mm(2) per% oxidation for LR5. A second group of three peptides were also selected showing increased binding with LDL oxidation: GlyCystineCysCys (1.5x10(7) LDL/mm(2) per microg/ml), GlyLysLysCys-SH (10(7) LDL/mm(2) per microg/ml) and GlyLysLys-OH (5.6x10(7) LDL/mm(2) per microg/ml). The latter gave a linear increase in LDL binding with oxidation level (1.2x10(7) LDL particles/mm(2) per% oxidation). LDL concentration is around 2-3 mg/ml in plasma compared with the low detection levels with this method (1-10 microg/ml), allowing a strategy to be developed requiring the minimum sample volume and diluting with physiological buffer prior to assay. By using a comparative reading between LDL adsorption on surfaces from the first and second group of peptides (e.g. GlyCystineSerAspGlu and GlyLysLys-OH, respectively), LDL oxidation could be determined without knowledge of LDL concentration. Higher binding was seen on GlyCystineSerAspGlu than GlyLysLys-OH below 30% LDL oxidation, whereas above 30% oxidation the binding on the latter surface was greater. Simple correlation of this form could provide good tests for atherosclerosis risk.  相似文献   

5.
c-Myc, the protein product of protooncogene c-myc, functions in cell proliferation, differentiation, and neoplastic disease. In this study, recombinant c-Myc and Max proteins, encompassing DNA binding (basic region) and dimerization (helix-loop-helix/leucine zipper) domain of human origin, were expressed in bacteria as Myc87 and Max85. Myc87 was purified under denatured conditions and was renatured again. The dissociation constant for the protein dimers and for dimer/DNA complexes were not detectable by isothermal titration calorimetry because of the low degree of solubility of Myc87 and Max85. Therefore, we set up equations which were used to determine the dissociation constants from the proportion of protein-DNA complexes. The dimer dissociation constants in TBS were 5.90(+/-0.54)x10(-7)M for Max85/Max85 homodimer, 6.85(+/-0.25)x10(-3)M for Myc87/Myc87 homodimer, and 2.55(+/-0.29)x10(-8)M for Myc87/Max85 heterodimer, and the DNA-binding dissociation constants in TBS were 1.33(+/-0.21)x10(-9)M for Max85/Max85/DNA, 2.27(+/-0.08)x10(-12)M for Myc87/Myc87/DNA, and 4.43(+/-0.37)x10(-10)M for Myc87/Max85/DNA. In addition, we revealed that linoleic acid which is known as an inhibitor for the formation of Max/Max/DNA complex reduced the affinity of Max homodimer for DNA. This result indicates that linoleic acid may bind to the DNA-binding region of Max homodimer.  相似文献   

6.
Boraston AB  Chiu P  Warren RA  Kilburn DG 《Biochemistry》2000,39(36):11129-11136
The C-terminal carbohydrate-binding module (CBM17) from Clostridium cellulovorans cellulase 5A is a beta-1,4-glucan binding module with a preference for soluble chains. CBM17 binds to phosphoric acid swollen Avicel (PASA) and Avicel with association constants of 2.9 (+/-0.2) x 10(5) and 1.6 (+/-0.2) x 10(5) M(-1), respectively. The capacity values for PASA and Avicel were 11.9 and 0.4 micromol/g of cellulose, respectively. CBM17 did not bind to crystalline cellulose. CBM17 bound tightly to soluble barley beta-glucan and the derivatized celluloses HEC, EHEC, and CMC. The association constants for binding to barley beta-glucan, HEC, and EHEC were approximately 2.0 x 10(5) M(-1). Significant binding affinities were found for cello-oligosaccharides greater than three glucose units in length. The affinities for cellotriose, cellotetraose, cellopentaose, and cellohexaose were 1.2 (+/-0.3) x 10(3), 4.3 (+/-0.4) x 10(3), 3.8 (+/-0.1) x 10(4), and 1.5 (+/-0.0) x 10(5) M(-1), respectively. Fluorescence quenching studies and N-bromosuccinimide modification indicate the participation of tryptophan residues in ligand binding. The possible architecture of the ligand-binding site is discussed in terms of its binding specificity, affinity, and the participation of tryptophan residues.  相似文献   

7.
Phosphatases, C4 and C8 esterases, leucine and valine aminopeptidases, N-acetyl-beta-glucosaminidase, beta-glucosidase, beta-galactosidase and beta-glucuronidase were detected in extracts of the parasitic mite Psoroptes cuniculi. Lipase, trypsin-like and chymotrypsin-like activities were not present. Haemoglobin was hydrolysed by a detergent-soluble fraction of the mite extracts with a maximum hydrolysis between pH 3 and 5. Acid proteinase activity was greater against haemoglobin than bovine serum albumin. Inhibitors of cysteine, serine and metallo-proteinases failed to inhibit the hydrolysis of H-Pro-Thr-Glu-Phe-Phe(NO2)-Arg-Leu-OH while pepstatin A inhibited its hydrolysis in a dose-dependent manner (IC50 8.02 x 10(-11) M (+/- 0.30 x 10(-11). Thermal inactivation of the proteolytic activity followed an exponential decay pattern. Typical K(m) and Vmax values were 7.2 x 10(-5) (+/- 0.7 x 10(-5) M-1 and 1.13 x 10(-3) (+/- 0.05 x 10(-3) OD unit-1 min-1 respectively. Acid proteinase activity eluted from a size exclusion column in a single, major peak representing a molecular weight range of 21-24.5 kDa. The major endoproteinase of P. cuniculi therefore appears to be a cathepsin D-like aspartic proteinase.  相似文献   

8.
1. A procedure is described for determining the affinity constant K(a) and the phosphorylation constant k(p) for the inhibition by di-isopropyl phosphorofluoridate of erythrocyte acetylcholinesterase and serum cholinesterase. The procedure depends on the use of a specially designed reaction vessel with which incubation times as short as 1.2sec. could be obtained at any convenient temperature. 2. The K(a) of acetylcholinesterase decreased from 1.58 (+/-0.22)x10(-3)m at 5 degrees to 1.17 (+/-0.10)x10(-3)m at 25 degrees and the associated change in enthalpy was 2980 cal. 3. The k(p) of acetylcholinesterase increased from 11.9 (+/-0.7)min.(-1) at 5 degrees to 40.7 (+/-1.4)min.(-1) at 25 degrees , indicating an activational energy of 9600 cal. The change in entropy associated with K(a) was 23.5 cal. degree(-1) at 25 degrees . 4. At 5 degrees , the K(a) and k(p) of serum cholinesterase were 9.95 (+/-1.10)x10(-6)m and 11.2 (+/-0.63)min.(-1) respectively. 5. The 150-fold difference in the inhibitory power of di-isopropyl phosphorofluoridate for the two cholinesterases was attributed entirely to differences in affinity.  相似文献   

9.
Cross-linked rabbit muscle phosphofructokinase in the active tetrameric and octameric state was studied in solution by hydrodynamic methods and small angle x-ray scattering techniques. The translational diffusion coefficients were determined by means of inelastic light scattering and were found to be 3.60 (+/- 0.02) x 10(-7) cm2 . s-1 for the tetramer and 2.54 (+/- 0.15) x 10(-7) cm2 . s-1 for the octamer. From small angle x-ray scattering measurements the radius of gyration, the specific inner surface area, and the volume were determined for both enzyme forms, revealing that the octameric cross-linked form is approximately spherical, with a diameter of 120.0 A, whereas the tetrameric form is asymmetric having an axial ratio of 2. By comparison of the scattering curves with triaxial geometric bodies which are equivalent in scattering, the tetrameric enzyme is described as a rectangular prism, with overall dimensions of A = 131.0 A, B = 131.0 A, and C = 65.0 A, and the octameric form as that of a cube with A = B = C = 120.0 A. The shape of the protomer, having a radius of gyration of 24.8 A, in the tetramer and octamer is similar to that for the native tetramer at pH 10 in the presence of 5 mM fructose 6-phosphate or 15 mM fructose 1,6-bis-phosphate. From the different shapes of the scattering curves of the native phosphofructokinase at pH 7.5 in the presence of 15 mM ATP and of the cross-linked tetramer or octamer, it can be inferred that the shapes of the protomers are different: in the presence of ATP the protomers are elongated, having an axial ratio of 1.8 to 2.0; the cross-linked state reveals a spherical protomer of radius 33.0 A, similar to that of the native enzyme at pH 7.5 in the presence of fructose 6-phosphate or fructose 1,6-bisphosphate.  相似文献   

10.
A high-performance liquid chromatographic (HPLC) method with electrochemical detection and solid-phase extraction (SPE) using cartridges of weak cation-exchange capacity as the primary retention mechanism is described for the separation and determination of methylnaltrexone (MNTX) in small clinical samples of plasma or urine. The procedure was performed using a Phenomenex Prodigy ODS-2, 5 microm, 150x3.2 mm analytical column and 50 mM potassium acetate buffer, with 11% methanol as organic modifier at pH* 4.5 at a flow-rate of 0.5 ml/min. The detection potential was 700 mV. The six-point standard calibration curves were linear over three consecutive days in the range from 2 to 100 ng/ml. The average goodness of fit (r) was 0.9993. The lower limit of detection (LOD) and limit of quantification (LOQ) were found to be 2.0 and 5.0 ng/ml, respectively. At the LOQ, the coefficient of variation for the entire method was 8.0% and the accuracy was 10.0% (n = 10). Recovery of the drug from plasma was in the region of 94%. The method was applied to a pharmacokinetics study of methylnaltrexone after subcutaneous administration and in numerous assays of analytes in blood plasma and urine. The pharmacokinetics parameters for a single dose of 0.1 or 0.3 mg/kg in plasma were C(max) = 110 (+/-55) and 287 (+/-101) ng/ml and t(max) = 16.7 (+/-10.8) and 20.0 (+/-9.5) min, respectively. The method is simple, yet sensitive for the detection and determination of methylnaltrexone in biological samples at the level of the physiological response.  相似文献   

11.
The effects of high and fluctuating pressure up to 220 bar on microbial growth and activity were determined in a pilot-scale water hydraulic system. An increase in the pipeline pressure from 70 to 220 bar decreased the total and the viable cell number in the pressure medium from 2.2(+/-0.5)x10(5) to 4.9(+/-1.5)x10(4) cells/ml and from 5.7(+/-2.8)x10(4) to 1.3(+/-0.7)x10(4) cfu/ml, respectively. Microbial attachment in the non-pressurised tank of the hydraulic system increased with increasing pipeline pressure [from 1.0(+/-0.3) to 3.8(+/-2.7)x10(5) cells/cm(2) on stainless steel]. The phosphatase, aminopeptidase and beta-glucosidase activities in the pressurised medium were between 0.02 and 1.4 micromol/lh ( V(max)) and decreased in response to increasing pipeline pressure. The alpha-glucosidase activity was detected only at 70 bar and the glucuronidase activity only occasionally. Based on principal component and cluster analyses, both the pressure applied and the original filling water quality affected substrate utilisation patterns. This study demonstrated the capability of freshwater bacteria to tolerate high and fluctuating pressure in a technical water system. Microbial survival was due to attachment and growth on the surfaces of the non-pressurised components and the nutrient flux released by cell lysis in the pressurised components. In summary, high pressures in water hydraulic systems do not prevent potential microbiologically related operational problems.  相似文献   

12.
The functional properties of the recombinant C-terminal dimerization domain of the Pseudomonas aeruginosa Fur (ferric uptake regulator) protein expressed in and purified from Escherichia coli have been evaluated. Sedimentation velocity measurements demonstrate that this domain is dimeric, and the UV CD spectrum is consistent with a secondary structure similar to that observed for the corresponding region of the crystallographically characterized wild-type protein. The thermal stability of the domain as determined by CD spectroscopy decreases significantly as pH is increased and increases significantly as metal ions are added. Potentiometric titrations (pH 6.5) establish that the domain possesses a high-affinity and a low-affinity binding site for metal ions. The high-affinity (sensory) binding site demonstrates association constants (K(A)) of 10(+/-7)x10(6), 5.7(+/-3)x10(6), 2.0(+/-2)x10(6) and 2.0(+/-3)x10(4) M(-1) for Ni2+, Zn2+, Co2+ and Mn2+ respectively, while the low-affinity (structural) site exhibits association constants of 1.3(+/-2)x10(6), 3.2(+/-2)x10(4), 1.76(+/-1)x10(5) and 1.5(+/-2)x10(3) M(-1) respectively for the same metal ions (pH 6.5, 300 mM NaCl, 25 degrees C). The stability of metal ion binding to the sensory site follows the Irving-Williams order, while metal ion binding to the partial sensory site present in the domain does not. Fluorescence experiments indicate that the quenching resulting from binding of Co2+ is reversed by subsequent titration with Zn2+. We conclude that the domain is a reasonable model for many properties of the full-length protein and is amenable to some analyses that the limited solubility of the full-length protein prevents.  相似文献   

13.
From 20 patients with solid tumors or acute nonlymphocytic leukemia in remission, hemopoietic progenitor cells were taken and stored in liquid nitrogen, for use in autologous bone marrow transplantation. Bone marrow aspiration resulted in a volume of 920(+/- 170) ml containing 16.8(+/-6.0) x 10(9) nucleated bone marrow cells and 7.2(+/-4.4) x 10(6) myeloid progenitor cells (CFUc). With use of the Haemonetics blood cell separator a progenitor cell-enriched fraction is obtained. This fraction is depleted of 90(+/-6)% of the erythrocytes and 59(+/-15)% of the neutrophils contained in the original. The original aspirate volume is reduced to one-fifth (21 +/- 3%) while containing 88(+/-38)% of the original CFUc's and 52(+/-11)% of the nucleated bone marrow cells. This technique of bone marrow enrichment has the advantage of a minimum of open-air contact, being independent of extensive laboratory facilities and manpower. The enriched fraction is frozen in autologous plasma and a final concentration of 10% (v/v) DMSO, using a program-controlled freezer (L'Air Liquide). Materials are stored at liquid nitrogen temperature in bags (Gambro) and test vials. Total CFUc recovery in test vials after thawing was 81(+/-32)%.  相似文献   

14.
Payne JC  Rous BW  Tenderholt AL  Godwin HA 《Biochemistry》2003,42(48):14214-14224
Zinc binding to the two Cys(4) sites present in the DNA-binding domain (DBD) of nuclear hormone receptor proteins is required for proper folding of the domain and for protein activity. By utilizing Co(2+) as a spectroscopic probe, we have characterized the metal-binding properties of the two Cys(4) structural zinc-binding sites found in the DBD of human estrogen receptor alpha (hERalpha-DBD) and rat glucocorticoid receptor (GR-DBD). The binding affinity of Co(2+) to the two proteins was determined relative to the binding affinity of Co(2+) to the zinc finger consensus peptide, CP-1. Using the known dissociation constant of Co(2+) from CP-1, the dissociation constants of cobalt from hERalpha-DBD were calculated: K(d1)(Co) = 2.2 (+/- 1.0) x 10(-7) M and K(d2)(Co) = 6.1 (+/- 1.5) x 10(-7) M. Similarly, the dissociation constants of Co(2+) from GR-DBD were calculated: K(d1)(Co) = 4.1 (+/- 0.6) x 10(-7) M and K(d2)(Co) = 1.7 (+/- 0.3) x 10(-7) M. Metal-binding studies conducted in which Zn(2+) displaces Co(2+) from the metal-binding sites of hERalpha-DBD and GR-DBD indicate that Zn(2+) binds to each of the Cys(4) metal-binding sites approximately 3 orders of magnitude more tightly than Co(2+) does: the stoichiometric dissociation constants are K(d1)(Zn) = 1 (+/- 1) x 10(-10) M and K(d2)(Zn) = 5 (+/- 1) x 10(-10) M for hERalpha-DBD and K(d1)(Zn) = 2 (+/- 1) x 10(-10) M and K(d2)(Zn) = 3 (+/- 1) x 10(-10) M for GR-DBD. These affinities are comparable to those observed for most other naturally occurring structural zinc-binding sites. In contrast to the recent prediction by Low et. al. that zinc binding in these systems should be cooperative [Low, L. Y., Hernández, H., Robinson, C. V., O'Brien, R., Grossmann, J. G., Ladbury, J. E., and Luisi, B. (2002) J. Mol. Biol. 319, 87-106], these data suggest that the zincs that bind to the two sites in the DBDs of hERalpha-DBD and GR-DBD do not interact.  相似文献   

15.
Electrochemical methods were used to activate MnIII and FeIII complexes of meso-tetrakis(N-methyl-4-pyridiniumyl)porphine (H2TMPyP) to cause cleavage of pBR322 DNA and to study their interaction with sonicated calf thymus DNA. Electrochemical reduction of MnIIITMPyP and FeIIITMPyP (at low concentrations) in the presence of O2 was required to activate these complexes. However, FeIIITMPyP at 1 x 10(-6) M produced DNA strand breakage without being electrochemically reduced. At low concentrations, FeIITMPyP was more efficient at cleaving DNA than MnIITMPyP. Reduction of O2 at a platinum electrode also produced some cleavage but to a much smaller extent. The oxidized form of MnIIITMPyP (charge 5+) has higher affinity for sonicated calf thymus (CT) DNA than the reduced form (charge 4+), as determined by the negative shift in E degrees' for the voltammetric wave in the presence of DNA. Both forms of FeIIITMPyP (charge 4+) interact with DNA to about the same extent. Differential pulse voltammetry was used to determine binding constants (K) and binding-site sizes (s) of the interaction of these metalloporphyrins with sonicated CT DNA. The data were analyzed assuming both mobile and static equilibria. MnIIITMPyP binds to DNA (5 mM Tris, 50 mM NaCl, pH 7) with K = 5 (+/- 2) x 10(6) M-1, s = 3 bp (mobile) or K = 3.6 (+/- 0.3) x 10(6) M-1, s = 4 bp (static). FeIIITMPyP at that ionic strength caused DNA precipitation. At higher ionic strength (0.1 M Tris, 0.1 M NaCl, pH 7), FeIIITMPyP associates to DNA with K = 4.4 (+/- 0.2) x 10(4) M-1, s = 5 bp (mobile) or K = 1.9 (+/- 0.1) x 10(4) M-1, s = 6 bp (static).  相似文献   

16.
H Yu  N Soong    W F Anderson 《Journal of virology》1995,69(10):6557-6562
A quantitative analysis of the binding kinetics of intact Moloney murine leukemia retrovirus (MoMuLV) particles with NIH 3T3 cells was performed with an immunofluorescence flow cytometry assay. The virus-cell binding equilibrium dissociation constant (KD), expressed in terms of virus particle concentration, was measured to be 8.5 (+/- 6.4) x 10(-12) M at 4 degrees C and was three- to sixfold lower at temperatures above 15 degrees C. The KD of virus binding is about 1,000-fold lower than the KD of purified MoMuLV envelope. The association rate constant was determined to be 2.5 (+/- 0.9) x 10(9) M-1 min-1 at 4 degrees C and was 5- to 10-fold higher at temperatures above 15 degrees C. The apparent dissociation rate constant at 4 degrees C was 1.1 (+/- 0.4) x 10(-3) min-1 and was doubled for every 10 degrees C increase in temperature over the range tested (15 to 37 degrees C).  相似文献   

17.
We have previously used NotI as the marker enzyme (recognizing GCGGCCGC) in a genome scanning approach for detection of mutations induced in mouse spermatogonia and estimated the mutation induction rate as about 0.7 x 10(-5) per locus per Gy. To see whether different parts of the genome have different sensitivities for mutation induction, we used AflII (recognizing CTTAAG) as the marker enzyme in the present study. After the screening of 1,120 spots in each mouse offspring, we found five mutations among 92,655 spots from the unirradiated paternal genome, five mutations among 218,411 spots from the unirradiated maternal genome, and 13 mutations among 92,789 spots from 5 Gy-exposed paternal genome. Among the 23 mutations, 11 involved mouse satellite DNA sequences (AT-rich), and the remaining 12 mutations also involved AT-rich but non-satellite sequences. Both types of sequences were found as multiple, similar-sequence blocks in the genome. Counting each member of cluster mutations separately and excluding results on one hypermutable spot, the spontaneous mutation rates were estimated as 3.2 (+/- 1.9) x 10(-5) and 2.3 (+/- 1.0) x 10(-5) per locus per generation in the male and female genomes, respectively, and the mutation induction rate as 1.1 (+/- 1.2) x 10(-5) per locus per Gy. The induction rate would be reduced to 0.9 x 10(-5) per locus per Gy if satellite sequence mutations were excluded from this analysis. The results indicate that mutation induction rates do not largely differ between GC-rich and AT-rich regions: 1 x 10(-5) per locus per Gy or less, which is close to 1.08 x 10(-5) per locus per Gy, the current estimate for the mean mutation induction rate in mice.  相似文献   

18.
Placental neurokinin B (NKB) was recently identified as the causative agent in preeclampsia, a condition characterized by increased maternal and feto-placental vascular resistance. We hypothesized that NKB should constrict placental resistance vessels. Placentas were obtained from normotensive pregnancies. Immediately after delivery, stem villous arteries (300 microm diameter, 1.2 mm long) were dissected from macroscopically normal tissue in cold HEPES-physiological salt solution (PSS), mounted on a wire myograph system, and bathed in HEPES-PSS at 37 degrees C. After determination of the passive-tension internal circumference characteristics, the arteries were set to 90% of the internal circumference they would have under a normal physiological transmural pressure. Cumulative concentration-response curves were constructed for NKB (1 x 10(-12) to 1 x 10(-5) mol/l). Since there was no constrictive response to NKB, cumulative constrictive concentration-response curves were constructed to the thromboxane A(2) mimetic U46619 (1 x 10(-9) to 1 x 10(-5) mol/l). The vessels were then pre-constricted to 80% of maximal response and exposed to cumulative concentrations of NKB (1 x 10(-12) to 1 x 10(-6) mol/l). NKB caused a concentration-dependent relaxation (Maximal response NKB, 51+/-5%, n=5; time control, 12+/-6%, n=4; P<0.05). Removal of the endothelium did not alter the vasodilatory response to NKB. We conclude that, contrary to our hypothesis, NKB causes an endothelium-independent relaxation of the placental resistance vessels. We propose that NKB plays a role in the maintenance of high placental blood flow in normal pregnancy.  相似文献   

19.
Bovine pancreatic ribonuclease A (RNase A) catalyzes the cleavage of P-O5' bonds in RNA on the 3' side of pyrimidine to form cyclic 2', 5'-phosphates. It has several high affinity binding sites that make it possible target for many organic and inorganic molecules. Ligand binding to RNase A can alter protein secondary structure and its catalytic activity. In this review, the effects of several drugs such as AZT (anti-AIDS), cis-Pt (antitumor), aspirin (anti-inflammatory), and vitamin C (antioxidant) on the stability and conformation of RNase A in vitro are compared. The results of UV-visible, FTIR, and CD spectroscopic analysis of RNase complexes with aspirin, AZT, cis-Pt, and vitamin C at physiological conditions are discussed here. Spectroscopic results showed one major binding for each drug-RNase adduct with KAZT=5.29 (+/-1.6)x10(4) M(-1), Kaspirin=3.57 (+/-1.4)x10(4) M(-1), Kcis-Pt=5.66 (+/-1.9)x10(3) M(-1), and Kascorbate=3.50 (+/-1.5)x10(3) M(-1). Major protein unfolding occurred with reduction of alpha-helix from 29% (free protein) to 20% and increase of beta-sheet from 39% (free protein) to 45% in the aspirin-, ascorbate-, and cis-Pt-RNase complexes, while minor increase of alpha-helix was observed for AZT-RNase adduct.  相似文献   

20.
Depending on solution conditions, beta-lactoglobulin can exist in one of its six pH-dependent structural states. We have characterized the acid and basic-induced conformational transitions between these structural states over the pH range of pH 1 to pH 13. To this end, we have employed high-precision ultrasonic and densimetric measurements coupled with fluorescence and CD spectroscopic data. Our combined spectroscopic and volumetric results have revealed five pH-induced transitions of beta-lactoglobulin between pH 1 and pH 13. The first transition starts at pH 2 and is not completed even at pH 1, our lowest experimental pH. This transition is followed by the dimer-to-monomer transition of beta-lactoglobulin between pH 2.5 and pH 4. The dimer-to-monomer transition is accompanied by decreases in volume, v degrees (-0.008(+/-0.003) cm3 x g(-1)), and adiabatic compressibility, k degrees (S) (-(0.7(+/-0.4))x10(-6) cm3 x g(-1) x bar(-1)). We interpret the observed changes in volume and compressibility associated with the dimer-to-monomer transition of beta-lactoglobulin, in conjunction with X-ray crystallographic data, as suggesting a 7 % increase in protein hydration, with the hydration changes being localized in the area of contact between the two monomeric subunits. The so-called N-to-Q transition of beta-lactoglobulin occurs between pH 4.5 and pH 6 and is accompanied by increases in volume, v degrees (0.004(+/-0.003) cm3 x g(-1)), and compressibility, k degrees (S) ((0.7(+/-0.4))x10(-6) cm3 x g(-1) x bar(-1)). The Tanford transition of beta-lactoglobulin is centered at pH 7.5 and is accompanied by a decrease in volume, v degrees (-0.006(+/-0.003) cm3 x g(-1)), and an increase in compressibility, k degrees (S) ((1.5(+/-0.5))x10(-6) cm3 x g(-1) x bar(-1)). Based on these volumetric results, we propose that the Tanford transition is accompanied by a 5 to 10 % increase in the protein hydration and a loosening of the interior packing of beta-lactoglobulin as reflected in a 12 % increase in its intrinsic compressibility. Finally, above pH 9, the protein undergoes irreversible base-induced unfolding which is accompanied by decreases in v degrees (-0.014(+/-0.003) cm3 x g(-1)) and k degrees (S) (-(7.0(+/-0.5))x10(-6) cm3 x g(-1) x bar(-1)). Combining these results with our CD spectroscopic data, we propose that, in the base-induced unfolded state of beta-lactoglobulin, only 80 % of the surface area of the fully unfolded conformation is exposed to the solvent. Thus, in so far as solvent exposure is concerned, the base-induced unfolded states of beta-lactoglobulin retains some order, with 20 % of its amino acid residues remaining solvent inaccessible.  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