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.
7-Alkylguanosine and 7-alkyldeoxyguanosine were prepared from phosphoramide mustard and nitrogen mustard in nonaqueous conditions. The guanosine products were N-(2-chloroethyl)-N-[2-(7-guanosinyl)ethyl] phosphorodiamidic acid, and N-(2-chloroethyl)-N-[2-(7-guanosinyl)ethyl]methylamine, respectively. These were also formed in aqueous reactions but they rapidly underwent secondary reactions. The half-life of the phosphoramide mustard-guanosine adduct was 3.1 h (37 degrees C, pH 7.4) and that of the nitrogen mustard adduct 1 h (25 degrees C, pH 7.4), as determined by HPLC. The half-lives of the respective adducts for imidazole ring-opening were 9.5 h and 0.78 h (37 degrees C, pH 7.4). The respective deoxyguanosine derivatives depurinated with half-lives of 8.5 h and 1.6 h (25 degrees C, pH 4.2). The mustard adducts are notably more labile than simple alkyl substituted guanosines and deoxyguanosines.  相似文献   

2.
When HeLa S3 cells were subjected to 45 degrees C hyperthermia, DNA lesions were detected by the use of the alkaline unwinding/hydroxylapatite method. The number of lesions formed was not affected when the cells were made thermotolerant by either an acute (15 min 44 degrees C + 5 h 37 degrees C) or a chronic (5 h 42 degrees C) pretreatment before 45 degrees C hyperthermia. The presence of 10 mM procaine (heat sensitizer) or 0.5 M erythritol (heat protector) during hyperthermia also had no effect on the rate of formation of heat-induced alkali labile DNA lesions. These observations do not support a concept where DNA lesions are considered to be the ultimate cause of hyperthermic cell killing. Both drugs, however, influenced the rate of repair of radiation-induced strand breaks when present during preirradiation heat treatment. We conclude that the initial number of heat-induced alkali labile DNA lesions is not directly related to cell survival. It cannot be excluded, however, that differences in posthyperthermic repair of these lesions may lead to a positive correlation between residual DNA damage and survival after the different experimental conditions.  相似文献   

3.
Identification of the thiol ester linked lipids in apolipoprotein B   总被引:4,自引:0,他引:4  
G Huang  D M Lee  S Singh 《Biochemistry》1988,27(5):1395-1400
Human plasma low-density lipoproteins of 1.032-1.043 g/mL density were totally delipidized. The reduced and carboxymethylated apolipoprotein B was incubated with 50 mM [14C]methylamine at pH 8.5 at 30 degrees C. Covalent incorporation of [14C]methylamine was observed with concomitant generation of new sulfhydryl groups, which could be blocked with [3H]- or [14C]iodoacetic acid. One type of the [14C]methylamine-modified products was separated from the protein and was found to be lipid in nature. Its Rf on thin-layer chromatography (TLC) was similar to that of the synthetic N-methyl fatty acyl amides. After purification with TLC and transesterification in 3 N methanolic HCl, methyl esters of C16 and C18 fatty acids at 1:1 ratio were identified by gas-liquid chromatography. The transesterification method was verified with the known N-methyl fatty acyl amides. These results suggest the presence of labile thiol ester linked palmitate and stearate in apolipoprotein B. Under mild alkaline conditions, the thiol ester bonds are broken by methylamine and form N-methyl fatty acyl amides and release new-SH groups. Intramolecular thiol ester bonds linked between cysteine side chains and acidic amino acid residues were also found present, which will be reported separately.  相似文献   

4.
Growth of temperature-sensitive mutant Bacillus cereus T JS22-C occurred normally at the restrictive temperature (37 degrees C), but sporulation was blocked at stage 0. The production of extracellular and intracellular proteases and of alkaline phosphatase occurred at 37 degrees C, but the expression of a functional tricarboxylic acid cycle did not. At the permissive temperature (26 degrees C), the mutant sporulated at a slightly lower frequency (60%) and at a lower rate than the parent strain. The oxidation of organic acids, which accumulate in the growth medium began at T0 in cultures of the parent strain but was delayed until about T3 in cultures of the mutant. Later events in sporulation were also delayed in the mutant by about 3 h. Experiments in which the temperature of growth was shifted from 37 to 26 degrees C or from 26 to 37 degrees C at various times showed that the temperature-sensitive event began approximately 1 h after the end of exponential growth and ended when the cells reached the end of stage II (septum formation). The absence of a functional tricarboxylic acid cycle in cells of the mutant grown at 37 degrees C or shifted from 26 to 37 degrees C before T1 did not appear to be due to a lesion in one of the structural genes of the tricarboxylic acid cycle but was more likely due to the inability of the cells to derepress the synthesis of some of the enzymes of that cycle.  相似文献   

5.
The mechanism of ethionine carcinogenesis and more generally the relationship between alkylation of nucleic acids by chemical carcinogens and oncogenesis still remain obscure. In the present study the rat liver tRNA ethylation by L-[ethyl-1-3H]ethionine was reinvestigated by examining in particular the highly radioactive 'pyrimidine-nucleotide-like' fraction found earlier in acid hydrolysates of hepatic tRNA from ethionine-treated rats. The following results were obtained: (1) ultraviolet-spectral and chromatographic analyses showed the presence of 1,7-diethylguanosine in this 'pyrimidine-nucleotide-like' fraction; (2) the dialkyl compound was recovered exclusively in the form of imidazole-ring-opened derivatives. When [1-14C]ethylnitrosourea was used as alkylating agent, the in vivo ethylation pattern of tRNA from various organs of rat showed an analogous radioactive 'pyrimidine-nucleotide-like' fraction as main radioactive product. On the contrary, tRNA ethylation pattern after in vitro reaction with [1-14C]ethylnitrosourea exhibited a main radioactivity peak (85% of the total radioactivity recovered) in coincidence of the chromatographic area of 1,7-diethylguanine. The 1,7-diethylguanosine moieties of tRNA were extremely labile both under physiological and alkaline conditions. The 1,7-diethylguanine-associated radioactivity was completely lost from [14C]ethyl-tRNA after only 7 h incubation at 37 degrees C and pH 7.3, while at pH 11.4 this process was preceded by the conversion of the 1,7-diethylguanosine residues into imidazole-ring-opened derivatives.  相似文献   

6.
Reannealed hybrid molecules of wild-type bacteriophage lambda DNA were prepared in aqueous solutions of formamide at a variety of NaCl concentrations at both room temperature ( 22 degrees C) and 37 degrees C. Treatment of the hybrid DNA molecules with the single-strand-specific nuclease S1 from Aspergillus oryzae followed by alkaline sucrose gradient sedimentation was used to monitor the extent and fidelity of hybridization. The optimal renaturation conditions at room temperature were found to be: 50% formamide, 35-55 mM NaCl and 10 mM Tris-HCl (pH 8.5) at 20-25 mug DNA/ml. Optimal conditions at 37 degrees C were: 32% formamide, 35-55 mM NaCl and 10 mM Tris-HCl (pH 8.5) at 20-25 mug DNA/ml. Under these conditions approximately 85-90% of the input single-stranded DNA (molecular weight 1.5 X 10(7)) was rendered S1-nuclease-resistant within 8 h at room temperature and 5 h at 37 degrees C. Neither Mg2+ nor spermidine appeared to have an effect on either the extent or fidelity of duplex formation. Experiments performed with excess enzyme and with lambda/lambda imm 434 heteroduplex hybrids suggested that the hybrid that the hybrid DNA molecules formed under optimal conditions contained no, or only short (less than 1%), mismatched regions.  相似文献   

7.
A cell line (HuL-1) derived from normal fetal human liver was adapted to grow continuously in a modified Eagle's minimum essential medium without serum or hormones. The population doubling time of this adapted cell line (HuL-1-317) was about 72 h and the modal number of chromosomes was 54. The morphology of HuL-1-317 cells was round in the absence of serum, but at 37 degrees C with the addition of serum (1-10%), the cells flattened. HuL-1-317 cells had a low level of alkaline phosphatase activity. However the enzyme activity was slightly enhanced by the combination of prednisolone, butyrate, dibutyryl cyclic adenosine monophosphate and a hypertonic concentration of NaCl after 3 days of incubation at 37 degrees C. The increase in alkaline phosphatase activity with the four agents was further amplified dose-dependently by the pretreatment of the cells with serum. The stimulatory effect of the serum was evident at concentrations as low as 1%, and was maximal at 20%. The half life of the effect of serum on alkaline phosphatase induction was 48 h at 37 degrees C. Serum alone could not enhance the enzyme activity without the four agents. The present results indicate that serum contributes to the regulation of alkaline phosphatase induction by the combination of prednisolone, butyrate, dibutyryl cyclic adenosine monophosphate and NaCl in fetal human liver cells (HuL-1-317).  相似文献   

8.
2',3'-O-(2,4,6-trinitrophenyl)-8-azido (TNP-8N3)-AMP, -ADP, and -ATP bind tightly to the Ca(2+)-ATPase of skeletal muscle sarcoplasmic reticulum and become covalently attached on irradiation at alkaline pH, concomitant with inactivation of ATPase activity (Seebregts, C. J., and McIntosh, D. B. (1989) J. Biol. Chem. 264, 2043-2052). The ATPase is derivatized to the extent of 2-3 nmol/mg protein (i.e. approximately 1/2 maximum phosphoenzyme levels) per irradiation period at equimolar concentrations of ATPase and nucleotide. Stability studies of the adduct formed at alkaline pH revealed that the linkage is labile, particularly if the protein is denatured by brief heat (60 degrees C) treatment (t1/2 = 4-8 h at 40 degrees C). Thermolysin digestion of derivatized vesicles resulted in the release of the majority of the TNP chromaphore as an unstable TNP-peptide adduct (t1/2 = 9 h at 25 degrees C) with the sequence FSRDR*SMS, where the missing residue is Lys-492 and is presumably that which is derivatized. The same peptide adduct, and in similar amounts, was isolated from the ATPase derivatized with either TNP-8N3-AMP or -ATP. Several lines of evidence, including the finding that ATP- and not acetyl phosphate- or Pi-dependent phosphorylation is blocked by derivatization, suggest that the lysyl residue is at the catalytic nucleotide binding site, but is not directly involved in phosphoryl transfer. Lys-492 and Phe-487, as well as neighboring Arg-476 and Lys-515 (labeled with fluorescein 5'-isothiocyanate), have all been highly conserved and probably contribute to a subdomain binding the purine and/or proximal phosphoryl groups of ATP.  相似文献   

9.
We have obtained biochemical and electron microscopic evidence of conformational changes at pH 8.0 and 37 degrees C in the coronavirus spike glycoprotein E2 (S). The importance of these changes is reflected in the loss of virus infectivity, the aggregation of virions, and increased virus-induced cell fusion at the same pH. Coronavirus (MHV-A59) infectivity is exquisitely sensitive to pH. The virus was quite stable at pH 6.0 and 37 degrees C (half-life, approximately 24 h) but was rapidly and irreversibly inactivated by brief treatment at pH 8.0 and 37 degrees C (half-life, approximately 30 min). Virions treated at pH 8.0 and 37 degrees C formed clumps and large aggregates. With virions treated at pH 8.0 and 37 degrees C, the amino-terminal peptide E2N (or S1) was released from virions and the remaining peptide, E2C (S2), was aggregated. Viral spikes isolated from detergent-treated virions also aggregated at pH 8.0 and 37 degrees C. Loss of virus infectivity and E2 (S) aggregation at pH 8.0 and 37 degrees C were markedly enhanced in the presence of dithiothreitol. On the basis of the effects of dithiothreitol on the reactions of the peplomer, we propose that release of E2N (S1) and aggregation of E2C (S2) may be triggered by rearrangement of intramolecular disulfide bonds. The aggregation of virions and the isolated E2 (S) glycoprotein at pH 8.0 and 37 degrees C or following treatment with guanidine and urea at pH 6.0 and 37 degrees C indicate that an irreversible conformational change has been induced in the peplomer glycoprotein by these conditions. It is interesting that coronavirus-induced cell fusion also occurred under mildly alkaline conditions and at 37 degrees C. Some enveloped viruses, including influenza viruses and alphaviruses, show conformational changes of spike glycoproteins at a low pH, which correlates with fusion and penetration of those viruses in acidified endocytic vesicles. For coronavirus MHV-A59, comparable conformational change of the spike glycoprotein E2 (S) and cell fusion occurred at a mildly alkaline condition, suggesting that coronavirus infection-penetration, like that of paramyxoviruses and lentiviruses, may occur at the plasma membrane, rather than within endocytic vesicles.  相似文献   

10.
Acetone-photosensitized UV irradiation of three thymine oligomers, d(TpT), d(TpTpT), and d(TpTpTpT), forms predominantly cis-syn cyclobutyl photodimers. C-18 reverse-phase high-performance liquid chromatography is used to purify the following positional isomers: d(TpT[p]T), d(T[p]TpT), d(TpTpT[p]T), d(TpT[p]TpT), d(T[p]TpTpT), and d(T[p]TpT[p]T), where T[p]T represents the cis-syn photodimer. Conformational properties of the cis-syn dimers and adjacent thymine nucleotides have been investigated in solution by using 1H, 13C, and 31P NMR spectroscopy. These studies show that (1) the photodimer conformation in longer oligothymidylates is similar to that in the dinucleoside monophosphate and (2) the cis-syn dimer induces alterations to a greater degree on the 5' side than on the 3' side of the photodimer. Specifically, the photodimer distorts the exocyclic bonds epsilon(C3'-O3') in Tp- and gamma(C5'-C4') in -pT[p]- on the 5' side and slightly alters the furanose equilibrium of the -pT nucleotide on the 3' side of the dimer.  相似文献   

11.
A breeding trial was conducted to evaluate the effect of in vitro storage time and temperature on fertilizing capacity of equine spermatozoa. Semen obtained from one stallion and diluted with skim milk-glucose extender was used to artificially inseminate 45 estrussynchronized mares. The mares were assigned to one of three treatment groups (15 mares per group): 1) insemination with fresh semen (collected within 0.5 h of use), 2) insemination with semen stored for 24 h at 20 degrees C or 3) insemination with semen stored for 24 h at 5 degrees C. The mares were inseminated daily during estrus, from the detection of a 35-mm follicle until ovulation, with 250 x 10(6) progressively motile spermatozoa (based on initial sperm motility of fresh semen). Semen samples (n = 35) were evaluated prior to insemination for percentages of total sperm motility (TSM), progressive sperm motility (PSM) and sperm velocity (SV). Single-cycle 15-d pregnancy rates. resulting from insemination with fresh semen, from fresh semen stored for 24 h at 20 degrees C or from semen stored for 24 h at 5 degrees C were the same (11 15 ; 73%). Mean diameters (mm) of 15-d embryonic vesicles were not different (P>0.05) among these three treatment groups (21.5 +/- 2.9, 19.6 +/- 2.6 and 20.5 +/- 3.6, respectively). Ten pregnant mares were aborted on Day 15 of gestation for use in another project. The pregnancy status of the 23 remaining pregnant mares was again determined at 35 to 40 d and 55 to 60 d of gestation. No pregnancy losses occurred during this time period. Mean TSM percentages were different (P<0.05) among the three groups: the fresh semen percentage was 89 +/- 2, semen stored for 24 h at 20 degrees C was 57 +/- 11 and semen stored for 24 h at 5 degrees C was 80 +/- 6. Similar differences were found for mean PSM and SV. Semen storage at either 20 or 5 degrees C for 24 h had no apparent effect on the fertilizing capacity of the extended semen samples; however, the reduction in all motility parameters tested was more dramatic in semen stored at 20 degrees C than that stored at 5 degrees C.  相似文献   

12.
The growth of Listeria monocytogenes and Yersinia enterocolitica colonies was studied on solid media at 4 and 8 degrees C under modified atmospheres (MAs) of 5% O2: 10% CO2: 85% N2 (MA1), 30% CO2: 70% N2 (MA2) and air (control). Colony radius, determined using computer image analysis, allowed specific growth rates (mu) and the time taken to detect bacterial colonies to be estimated, after colonies became visible. At 4 degrees C both MAs decreased the growth rates of L. monocytogenes by 1.5- and 3.0-fold under MA1 (mu = 0.02 h(-1)) and MA2 (mu = 0.01 h(-1)), respectively, as compared with the control (mu = 0.03 h(-1)). The time to detection of bacterial colonies was increased from 15 d (control) to 24 (MA1) and 29 d (MA2). At 8 degrees C MA2 decreased the growth rate by 1.5-fold (mu = 0.04 h(-1)) as compared with the control (mu = 0.06 h(-1)) and detection of colonies increased from 7 (control) to 9 d (MA2). At 4 degrees C both MAs decreased the growth rates of Y. enterocolitica by 1.5- and 2.5-fold under MA1 (mu = 0.03 h(-1)) and MA2 (mu = 0.02 h(-1)), respectively, as compared with the control (mu = 0.05 h(-1)). At 8 degrees C identical growth rates were obtained under MA1 and the control (mu = 0.07 h(-1)) whilst a decrease in the growth rate was obtained under MA2 (mu = 0.04 h(-1)). The detection of colonies varied from 6 (8 degrees C, aerobic) to 19 d (4 degrees C, MA2). Refrigerated modified atmosphere packaged foods should be maintained at 4 degrees C and below to ensure product safety.  相似文献   

13.
The new adduct N6-(2-carboxyethyl)adenine (N6-CEA) was prepared from 1-(2-carboxyethyl)adenine (1-CEA) by base catalyzed (Dimroth) rearrangement of 1-CEA. The structure of N6-CEA was assigned on the basis of UV spectra and electron impact and isobutane chemical ionization mass spectra. When the carcinogen beta-propiolactone was reacted in vitro with calf thymus DNA, 1-CEA but not N6-CEA was detected on paper chromatograms following acid hydrolysis of the DNA. When BPL-reacted single-stranded DNA was incubated at pH 11.7 (37 degrees C, 18 h) prior to acid hydrolysis, it was found that 1-CEA was completely converted to N6-CEA in DNA by Dimroth rearrangement, whereas no conversion occurred at pH 7.5. The extent of Dimroth rearrangement at various pHs and temperatures was determined for 1-CEA, 1-methyladenine (1-MeA), 1-(2-carboxyethyl)-deoxyadenosine-5'-monophosphoric acid (1-CEdAdo5'P) and the phosphodiester 5'-O-(2-carboxyethyl)phosphono-1-(2-carboxyethyl)deoxyadenosine (1-CE-Ado-5'-P-CE).  相似文献   

14.
AIMS: An investigation was carried out on the purification and characterization of an alkaline protease from Bacillus pumilus MK6-5. METHODS AND RESULTS: An alkalophilic Bacillus pumilus MK6-5 was grown in a laboratory fermenter containing 1% reverse osmosis concentrated cheese whey powder, 0.25% corn steep liquor, 1% glucose, 0.5% tryptone, 1% sodium citrate, 0.02% MgSO4.7H2O and 0.65% Na2CO3 at 35 degrees C and pH 9.6, agitation at 250 rev min(-1) and aeration of 1 vvm for 60 h. When the enzyme was purified using ammonium sulphate precipitation, ion exchange and gel filtration chromatographies, a 26.2% recovery of enzyme with 36.6-fold purification was recorded. The purified protease was found to be homogenous by SDS-PAGE with molecular mass estimate of 28 kDa. The enzyme was optimally active at pH 11.5 and temperature of 55-60 degrees C. The Km and kcat values observed with synthetic substrates at 37 degrees C and pH 8.0 were 1.1 mmol l(-1) and 624 s(-1) for Glu-Gly-Ala-Phe-pNA and 3.7 mmol l(-1) and 826 s(-1) for Glu-Ala-Ala-Ala-pNA, respectively. The kinetic data revealed that small aliphatic and aromatic residues were the preferred residues at the P1 position. Inhibition profile exhibited by PMSF suggested the B. pumilus protease to be an alkaline serine protease. CONCLUSIONS: Bacillus pumilus MK6-5 produced a calcium-dependent, thermostable alkaline serine protease. SIGNIFICANCE AND IMPACT OF THE STUDY: The thermostable alkaline protease from Bacillus pumilus MK6-5 will be extremely useful in ultrafiltration membrane cleaning due to its ability to work in broad pH and temperature ranges, and tolerance to detergents, unlike the mesophilic proteases which face these limitations.  相似文献   

15.
J S Taylor  D S Garrett  M P Cohrs 《Biochemistry》1988,27(19):7206-7215
The preparation, spectroscopic investigation, structure determination, conformational analysis, and modeling of the Dewar pyrimidinone photoproduct of thymidylyl-(3'----5')-thymidine, previously referred to as TpT3 [Johns, H. E., Pearson, M. L., LeBlanc, J. C., & Heilleiner, C. W. (1964) J. Mol. Biol. 9, 503-524], is described. TpT3 was prepared in quantitative yield by photolysis of an aqueous solution of the (6-4) photoproduct of TpT with Pyrex-filtered medium-pressure mercury arc light. TpT3 was analyzed by FAB MS, IR, UV, and 1H, 13C, and 31P NMR spectroscopy. The spectroscopic data led to the conclusion that TpT3 results from the photoisomerization of the pyrimidinone ring of the (6-4) product of TpT to its Dewar valence isomer. Torsion angle and interproton distance information derived from coupling constants and NOE data was used to constrain ring conformation searches by utilizing the SYBYL molecular modeling program subroutine SEARCH. Sets of angles derived from the ring search procedure were then used to construct structures whose geometries were optimized by the energy-minimization subroutine MAXIMIN. A two-state model for the solution-state structure of the Dewar photoproduct was chosen which was energetically sound, fit the experimental coupling constants with an RMS deviation of 1.15 Hz, and was consistent with the NOE data. The model for the Dewar photoproduct was compared to a model for the (6-4) photoproduct and the TpT subunits of the Dickerson dodecamer structure by a least-squares fitting procedure. It was concluded that the Dewar photoproduct more closely resembles a B-form TpT unit than does the (6-4) photoproduct.  相似文献   

16.
In the present study, we determined whether exposure of mammalian cells to 3.2-5.1 W/kg specific absorption rate (SAR) radiofrequency fields could induce DNA damage in murine C3H 10T(1/2) fibroblasts. Cell cultures were exposed to 847.74 MHz code-division multiple access (CDMA) and 835.62 frequency-division multiple access (FDMA) modulated radiations in radial transmission line (RTL) irradiators in which the temperature was regulated to 37.0 +/- 0.3 degrees C. Using the alkaline comet assay to measure DNA damage, we found no statistically significant differences in either comet moment or comet length between sham-exposed cells and those exposed for 2, 4 or 24 h to CDMA or FDMA radiations in either exponentially growing or plateau-phase cells. Further, a 4-h incubation after the 2-h exposure resulted in no significant changes in comet moment or comet length. Our results show that exposure of cultured C3H 10T(1/2) cells at 37 degrees C CDMA or FDMA at SAR values of up to 5.1 W/kg did not induce measurable DNA damage.  相似文献   

17.
An alkalophilic Aspergillus nidulans KK-99 produced an alkaline, thermostable xylanase (40 IU/ml) in a basal medium supplemented with wheat bran (2% w/v) and KNO3 (at 0.15% N) pH 10.0 and 37 degrees C. The partially purified xylanase was optimally active at pH 8.0 and 55 degrees C. The xylanase was stable in a broad pH range of 4.0-9.5 for 1 h at 55 degrees C, retaining more than 80% of its activity. The enzyme exhibited greater binding affinity for xylan from hardwood than from softwood. The xylanase activity was stimulated (+25%) by Na+ and Fe2+ and was strongly inhibited (maximum by 70%) by Tween-20, 40, 60, SDS, acetic anhydride, phenylmethane sulphonyl fluoride, Triton-X-100. The xylanase dose of 1.0 IU/g dry weight pulp gave optimum bleach boosting of Kraft pulp at pH 8.0 and temperature 55 degrees C for 3 h reaction time.  相似文献   

18.
A method was developed for direct microsequencing of N alpha-acetylated proteins electroblotted onto polyvinylidene difluoride membranes from polyacrylamide gels. N alpha-Acetylated proteins (greater than 32 pmol), including horse heart cytochrome c, five mutants of yeast cytochrome c, and bovine erythrocyte superoxide dismutase, were separated by SDS-PAGE and electroblotted onto polyvinylidene difluoride membranes. The portions of the membrane carrying the bands were cut out and treated with 0.5% polyvinylpyrrolidone in acetic acid solution at 37 degrees C for 30 min. The protein was digested on the membrane with 5-10 micrograms of trypsin at 37 degrees C for 24 h. During tryptic digestion, the resultant peptides were released from the membrane and the N-terminal peptide was efficiently deblocked with 50 mU of acylamino acid-releasing enzyme at 37 degrees C for 12 h. Picomole levels of the deblocked proteins could be sequenced directly by use of a gas-phase protein sequencer.  相似文献   

19.
The rates of intracellular protein degradation, of identically labelled populations of proteins, were compared in hepatocytes cultured at 37 degrees (on an adsorbed collagen layer) and in cells preserved on gelatin gels at 10 degrees C. The half-lives of the long-lived proteins were 35.4+/-8.6 h (N=4) and 692.9+/-216.9 h (N=4) respectively. Proteolysis was substantially decreased at 10 degrees C but the rate of decrease remained constant. Hepatocytes rapidly removed resorufin from the culture medium. The resorufin was not being conjugated or accumulated within the cells. Dicumarol, a potent inhibitor of quinone oxidoreductase, at high concentration (500 microm ) caused only a 72% decrease in the utilization of resorufin. The microsomal detoxifying enzyme, cytochrome P-450 1A1 remained at a constant level in the preserved hepatocyte monolayers. The results of this study strongly favour storing hepatocytes at 10 degrees C rather than at 4 degrees or 37 degrees C.  相似文献   

20.
The gene encoding alkaline phosphatase from the psychrotrophic bacterium Shewanella sp. SIB1 was cloned, sequenced, and overexpressed in Escherichia coli. The recombinant protein was purified and its enzymatic properties were compared with those of E. coli alkaline phosphatase (APase), which shows an amino acid sequence identity of 37%. The optimum temperature of SIB1 APase was 50 degrees C, lower than that of E. coli APase by 30 degrees C. The specific activity of SIB1 APase at 50 degrees C was 3.1 fold higher than that of E. coli APase at 80 degrees C. SIB1 APase lost activity with a half-life of 3.9 min at 70 degrees C, whereas E. coli APase lost activity with a half-life of >6 h even at 80 degrees C. Thus SIB1 APase is well adapted to low temperatures. Comparison of the amino acid sequences of SIB1 and E. coli APases suggests that decreases in electrostatic interactions and number of disulfide bonds are responsible for the cold-adaptation of SIB1 APase.  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