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 15 毫秒
1.
Stable isotope methodology was used to determine the kinetic behavior of apolipoprotein (apo) A-I within the triglyceride-rich lipoprotein (TRL) fraction and to compare TRL apoA-I kinetics with that of apoA-I in high density lipoprotein (HDL) and TRL apoB-48. Eight subjects (5 males and 3 females) over the age of 40 were placed on a baseline average American diet and after 6 weeks received a primed-constant infusion of [5,5,5-(2)H(3)]-l-leucine for 15 h while consuming small hourly meals of identical composition. HDL and TRL apoA-I and TRL apoB-48 tracer/tracee enrichment curves were obtained by gas chromatography;-mass spectrometry. Data were fitted to a compartmental model to determine the fractional secretion rates of apoA-I and apoB-48 within each lipoprotein fraction. Mean plasma apoA-I levels in TRL and HDL fractions were 0. 204 +/- 0.057 and 134 +/- 15 mg/dl, respectively. The mean fractional catabolic rate (FCR) of TRL apoA-I was 0.250 +/- 0.069 and HDL apoA-I was 0.239 +/- 0.054 pools/day, with mean estimated residence times (RT) of 4.27 and 4.37 days, respectively. The mean TRL apoB-48 FCR was 5.2 +/- 2.0 pools/day and the estimated mean RT was 5.1 +/- 1.8 h. Our results indicate that apoA-I is catabolized at a slower rate than apoB-48 within TRL, and that apoA-I within TRL and HDL fractions are catabolized at similar rates.  相似文献   

2.
Summary Apolipoprotein (apo) E, a major protein component of plasma lipoproteins, is a physiological ligand for the low density lipoprotein (LDL) receptor as well as for a specific apoE receptor; it is therefore an important modulator of lipoprotein metabolism. In this study we cloned and sequenced bovine apoE complementary DNA. Comparison of nucleotide substitution rates shows that apoE is less conservative than apoA-I and evolves about 30% faster than an average mammalian protein. Although apoE is not a conservative protein, several regions have been well conserved among all eight mammalian sequences now available. These include a 33-amino-acid block immediately upsteam from the third intron/exon junction and the LDL receptor binding region. We have also compared published apoC-I and apoC-II sequences. Both proteins are less conservative than apoE. In particular, apoC-I shows no well-conserved region except for a small region in the common 33-amino-acid block, suggesting that the function of apoC-I does not have stringent structural requirements. On the other hand, in apoC-II the region encoded by exon 4, which consists of the last 29 amino acids of the polypeptide, has been rather well conserved, probably because this region is important for the activation of lipoprotein lipase and chylomicron and very low density lipoprotein metabolism.  相似文献   

3.
Structure, evolution, and regulation of chicken apolipoprotein A-I   总被引:4,自引:0,他引:4  
A full-length cDNA clone for the precursor form of chicken liver apolipoprotein A-I (apoA-I) was isolated by antibody screening of a chicken liver cDNA library in the expression vector lambda gt11. The complete nucleotide sequence and predicted amino acid sequence of this clone is presented. The identity of the clone was confirmed by comparison with partial amino acid sequences for chicken apolipoprotein A-I. Chicken preproapolipoprotein A-1 consists of an 18-amino acid prepeptide, a 6-amino acid propeptide, and 240 amino acids of mature protein. The sequence of the protein is homologous to mammalian apoA-I and is highly internally repetitive, consisting largely of 11-amino acid repeats predicted to have an amphipathic alpha-helical structure. The sequence of the propeptide (Arg-Ser-Phe-Trp-Gln-His) differs in two positions from that of mammalian apoA-I. The mRNA for chicken apoA-I is about 1 kilobase in length and is expressed in a variety of tissues including liver, intestine, brain, adrenals, kidneys, heart, and muscle. This quantitative tissue distribution has been determined and is similar to that observed for mammalian apoE and different from that of mammalian apoA-I mRNA. This reinforces the concept that avian apoA-I performs functions analogous to those of mammalian apoE. Moreover, comparisons revealed sequences of chicken apoA-I similar to the region of mammalian apoE responsible for interaction with cellular receptors. Previous studies have demonstrated striking changes in the rates of synthesis of apoA-I in breast muscle during development and in optic nerve after retinal ablation. We now demonstrate that these changes are paralleled by changes in mRNA levels. ApoA-I mRNA levels increase approximately 50-fold in breast muscle between 14 days postconception and hatching and then decrease about 15-fold to adult levels. The levels of apoA-I mRNA increase about 3-fold in optic nerve following retinal ablation. ApoA-I mRNA is also found in the brain in the absence of nerve injury. This may indicate that locally synthesized apoA-I has a routine or housekeeping function in lipid metabolism in the central nervous system.  相似文献   

4.
Dog apolipoprotein (apo) C-I, A-I, and E cDNA clones were identified in a dog liver cDNA library in lambda gt10 by hybridization to synthetic oligonucleotide probes with the corresponding human DNA sequences. The longest clone for each apolipoprotein was completely sequenced. The apoC-I cDNA sequence predicts a protein of 62 residue mature peptide preceded by a 26 amino acid signal peptide. The apoA-I cDNA sequence predicts a 242 residue mature peptide, a 6 residue pro-segment, and an 18 residue signal peptide. The apoE cDNA, which lacks the signal peptide region, predicts a mature peptide of 291 amino acid residues. Slot blot hybridization of total RNA isolated from various dog tissues to dog apoC-I, A-I, and E cDNA probes indicates that apoC-I mRNA is detectable in liver only, apoA-I mRNA is present in liver and small intestine, though the concentration in the latter tissue is only approximately 15% of that in the liver, and apoE mRNA is present in multiple tissues including liver, jejunum, urinary bladder, ileum, colon, brain, kidney, spleen, pancreas, and testis with relative concentrations (%) of 100, 17.5, 7.5, 6.9, 5.9, 5.5, 5.0, 3.3, 1.0, and 1.0, respectively. These tissue distributions indicate that nascent lipoprotein particles produced in the dog small intestine would contain apoA-I and apoE but not apoC-I. The widespread tissue distribution of apoE mRNA indicates that like other mammals, peripheral synthesis of apoE contributes significantly to the total apoE pool in dog. We next compared the cDNA sequences among different vertebrate species for apoC-I (human and dog), A-I (human, rat, dog, rabbit and chicken), and E (human, rat, dog and rabbit) and calculated the rate of nucleotide substitution for each gene. Our results indicate that apoC-I has evolved rather rapidly and that on the whole, apoA-I is more conservative than apoE, contradictory to an earlier suggestion. ApoA-I is also more conservative than a region (residues 4204-4536) at the carboxyl-terminal portion, but less conservative than a region (residues 595-979) at the amino-terminal portion of apoB-100. Some regions in each of the apolipoproteins studied are better conserved than others and the rate of evolution of individual regions seems to be related to the stringency of functional requirements. Finally, we estimate that the human apoC-I pseudogene arose more than 35 million years ago, becoming nonfunctional soon after its formation.  相似文献   

5.
Using an antibody against chicken apolipoprotein (apo) A-I, we identified multiple cDNA clones for the protein in two intestinal cDNA libraries in λgtll. The complete nucleotide sequence of chicken apoA-I cDNA was determined. The sequence predicts a mature protein of 240 amino acids, a 6-amino acid propeptide and an 18-amino acid signal peptide. Using a 32P-cDNA probe, we detected the presence of apoA-I mRNA in 21 day old chicken intestine, liver, kidney, spleen, breast muscle and brain. The primary sequence of apoA-I contains numerous tandem repeats of 11 and 22 residues in a manner similar to the mammalian proteins. Our analysis of apoA-I sequences from human, rabbit, dog, rat, and chicken indicates that the rate of amino acid substitution is considerably faster in the rat lineage than in other mammalian lineages.  相似文献   

6.
The cDNA of mouse pancreatic mRNA has been cloned. After the library was screened with a rat ribonuclease cDNA probe, the positive clones were isolated and sequenced. There were no differences from the previously determined protein sequence. The mRNA codes for a preribonuclease of 149 amino acid residues including a signal peptide of 25 amino acids. The 3' noncoding region has a length of 260 bp, and the total mRNA length is approximately 940 bp. Comparison with the rat pancreatic ribonuclease sequence showed a high rate of nucleotide substitution. Within the coding region, nonsynonymous and synonymous substitution rates are 4.3 X 10(-9) and 15 X 10(-9) nucleotide substitutions/site/year, respectively. The latter value is one of the highest rates observed in the molecular evolution of mammalian nuclear genes. In the signal sequences the synonymous substitution rate is much lower and about the same as the nonsynonymous rate. Signal sequences of other mouse and rat proteins also exhibit little difference between synonymous and nonsynonymous rates. The sequences of rat and mouse pancreatic ribonuclease messengers were compared with those of bovine pancreatic, seminal, and brain ribonuclease. While the 3' noncoding regions of rat and mouse are very similar, as are those of the three bovine messengers, there is no significant similarity between both rodent and the three bovine messengers for the greater part of these regions. There is a duplication of approximately 50 nucleotides in the 3' noncoding region of the bovine messengers, with a region rich in A and C in between. The presence of this structural feature may be correlated with recent gene duplications that have occurred in the bovine genome.  相似文献   

7.
8.
The production rates of apolipoprotein (apo)B-100 in very low density lipoprotein and in low density lipoprotein and apolipoprotein A-I in high density lipoprotein were determined using a primed-constant infusion of [5,5,5,-2H3]leucine, [4,4,4,-2H3]valine, and [6,6-2H2,1,2-13C2]lysine. The three stable isotope-labeled amino acids were administered simultaneously to determine whether absolute production rates calculated using a stochastic model were independent of the tracer species utilized. Three normolipidemic adult males were studied in the constantly fed state over a 15-h period. The absolute production rates of very low density lipoprotein apoB-100 were 11.4 +/- 5.8 (leucine), 11.2 +/- 6.8 (valine), and 11.1 +/- 5.4 (lysine) mg per kg per day (mean +/- SDM). The absolute production rates for low density lipoprotein apoB-100 were 8.0 +/- 4.7 (leucine), 7.5 +/- 3.8 (valine), and 7.5 +/- 4.2 (lysine) mg per kg per day. The absolute production rates for high density lipoprotein apoA-I were 9.7 +/- 0.2 (leucine), 9.4 +/- 1.7 (valine, and 9.1 +/- 1.3 (lysine) mg per kg per day. There were no statistically significant differences in absolute synthetic rates of the three apolipoproteins when the plateau isotopic enrichment values of very low density lipoprotein apoB-100 were used to define the isotopic enrichment of the intracellular precursor pool. Our data indicate that deuterated leucine, valine, or lysine provided similar results when used for the determination of apoA-I and apoB-100 absolute production rates within plasma lipoproteins as part of a primed-constant infusion protocol.  相似文献   

9.
Nine hypercholesterolemic and hypertriglyceridemic subjects were enrolled in a randomized, placebo-controlled, double-blind, crossover study to test the effect of atorvastatin 20 mg/day and 80 mg/day on the kinetics of apolipoprotein B-100 (apoB-100) in triglyceride-rich lipoprotein (TRL), intermediate density lipoprotein (IDL), and LDL, of apoB-48 in TRL, and of apoA-I in HDL. Compared with placebo, atorvastatin 20 mg/day was associated with significant reductions in TRL, IDL, and LDL apoB-100 pool size as a result of significant increases in fractional catabolic rate (FCR) without changes in production rate (PR). Compared with the 20 mg/day dose, atorvastatin 80 mg/day caused a further significant reduction in the LDL apoB-100 pool size as a result of a further increase in FCR. ApoB-48 pool size was reduced significantly by both atorvastatin doses, and this reduction was associated with nonsignificant increases in FCR. The lathosterol-campesterol ratio was decreased by atorvastatin treatment, and changes in this ratio were inversely correlated with changes in TRL apoB-100 and apoB-48 PR. No significant effect on apoA-I kinetics was observed at either dose of atorvastatin. Our data indicate that atorvastatin reduces apoB-100- and apoB-48-containing lipoproteins by increasing their catabolism and has a dose-dependent effect on LDL apoB-100 kinetics. Atorvastatin-mediated changes in cholesterol homeostasis may contribute to apoB PR regulation.  相似文献   

10.
Zhu HL  Atkinson D 《Biochemistry》2007,46(6):1624-1634
Human apolipoprotein A-I (apoA-I) is the principle apolipoprotein of high-density lipoproteins that are critically involved in reverse cholesterol transport. The intrinsically flexibility of apoA-I has hindered studies of the structural and functional details of the protein. Our strategy is to study peptide models representing different regions of apoA-I. Our previous report on [1-44]apoA-I demonstrated that this N-terminal region is unstructured and folds into approximately 60% alpha-helix with a moderate lipid binding affinity. We now present details of the conformation and lipid interaction of a C-terminal 46-residue peptide, [198-243]apoA-I, encompassing putative helix repeats 10 and 9 and the second half of repeat 8 from the C-terminus of apoA-I. Far-ultraviolet circular dichroism spectra show that [198-243]apoA-I is also unfolded in aqueous solution. However, self-association induces approximately 50% alpha-helix in the peptide. The self-associated peptide exists mainly as a tetramer, as determined by native electrophoresis, cross-linking with glutaraldehyde, and unfolding data from circular dichroism (CD) and differential scanning calorimetry (DSC). In the presence of a number of lipid-mimicking detergents, above their CMC, approximately 60% alpha-helix was induced in the peptide. In contrast, SDS, an anionic lipid-mimicking detergent, induced helical folding in the peptide at a concentration of approximately 0.003% (approximately 100 microM), approximately 70-fold below its typical CMC (0.17-0.23% or 6-8 mM). Both monomeric and tetrameric peptide can solubilize dimyristoylphosphatidylcholine (DMPC) liposomes and fold into approximately 60% alpha-helix. Fractionation by density gradient ultracentrifugation and visualization by negative staining electromicroscopy demonstrated that the peptide binds to DMPC with a high affinity to form at least two sizes of relatively homogeneous discoidal HDL-like particles depending on the initial lipid:peptide ratio. The characteristics (lipid:peptide weight ratio, diameter, and density) of both complexes are similar to those of plasma A-I/DMPC complexes formed under similar conditions: small discoidal complexes (approximately 3:1 weight ratio, approximately 110 A, and approximately 1.10 g/cm3) formed at an initial 1:1 weight ratio and larger discoidal complexes (approximately 4.6:1 weight ratio, approximately 165 A, and approximately 1.085 g/cm3) formed at initial 4:1 weight ratio. The cross-linking data for the peptide on the complexes of two sizes is consistent with the calculated peptide numbers per particle. Compared to the approximately 100 A disk-like complex formed by the N-terminal peptide in which helical structure was insufficient to cover the disk edge by a single belt, the compositions of these two types of complexes formed by the C-terminal peptide are more consistent with a "double belt" model, similar to that proposed for full-length apoA-I. Thus, our data provide direct evidence that this C-terminal region of apoA-I is responsible for the self-association of apoA-I, and this C-terminal peptide model can mimic the interaction with the phospholipid of plasma apoA-I to form two sizes of homogeneous discoidal complexes and thus may be responsible for apoA-I function in the formation and maintenance of HDL subspecies in plasma.  相似文献   

11.
We have studied apolipoprotein synthesis, intracellular modification and secretion by primary adult rat hepatocyte cultures using continuous pulse or pulse chase labeling with [35S]methionine, immunoprecipitation and two-dimensional isoelectric focusing/polyacrylamide gel electrophoresis. The flotation properties of the newly secreted apolipoproteins were studied by discontinuous density gradient ultracentrifugation and one- and two-dimensional polyacrylamide gel electrophoresis. These studies showed that rat hepatocyte apoE is modified intracellularly to produce minor isoproteins that differ in size and charge. One of these minor isoproteins represents a monosialated apoE form (apoE3s1). Similarly, apoCIII is modified intracellularly to produce a disialated apoCIII form (apoCIIIs2), whereas newly synthesized apoA-I and apoA-IV are not glycosylated and overlap on two-dimensional gels with the proapoA-I and the plasma apoA-IV form, respectively. Both unmodified and modified apolipoproteins are secreted into the medium. Separation of secreted apolipoproteins by density gradient ultracentrifugation has shown that 50% of apoE, 80% of apoA-I, and more than 90% of apoA-IV and apoCIII are secreted in a lipid-poor form, whereas apoB-100 and apoB-48 are 100% associated with lipids. ApoB-100 floats in the VLDL and IDL regions, whereas apoB-48 is found in all lipoprotein fractions. ApoE and small amounts of apoA-I, apoA-IV and apoCIII float in the HDL region. Small amounts of apoE and apoCIII are also found in the VLDL and IDL regions, and apoE in the LDL region. Ultracentrifugation of nascent lipoproteins in the presence of rat serum promoted flotation of apoA-I and apoA-IV in the HDL fraction and resulted in increased flotation and distribution of apoE and apoCs in VLDL, IDL and LDL regions. These observations are consistent with the hypothesis that intracellular assembly of lipoproteins involves apoB-48 and apoB-100 forms, whereas a large portion of apoA-I, apoCIII and apoA-IV can be secreted in a lipid-poor form, which associates extracellularly with preexisting lipoproteins.  相似文献   

12.
The monolayer system was employed to investigate the relative affinities of apolipoproteins A-I and A-II for the lipid/water interface. The adsorption of reductively 14C-methylated apolipoproteins to phospholipid monolayers spread at the air/water interface was determined by monitoring the surface pressure of the mixed monolayer and the surface concentration of the apoprotein. ApoA-II has a higher affinity than apoA-I for lipid monolayers; for a given initial surface pressure, apoA-II adsorbs more than apoA-I to monolayers of egg phosphatidylcholine (PC), distearoyl-PC and human high-density lipoprotein (HDL3) surface lipids. Comparison of the molecular packing of apolipoproteins A-I and A-II suggests that apoA-II adopts a more condensed conformation at the lipid/water interface compared to apoA-I. The ability of apoA-II to displace apoA-I from egg PC and HDL3 surface lipid monolayers was studied by following the adsorption and desorption of the reductively 14C-methylated apolipoproteins. At saturating subphase concentrations of the apoproteins (3.10(-5) g/100 ml), two molecules of apoA-II absorbed for each molecule of apoA-I displaced. This displacement was accompanied by an increase in surface pressure. An identical stoichiometry for the displacement of apoA-I from HDL particles by apoA-II has been reported by others. At low subphase concentrations of apoproteins (5.10(-6) g/100 ml), the apoA-I/lipid monolayer was not fully compressed and could accommodate the adsorbing apoA-II molecules without displacement of apoA-I molecules. ApoA-I molecules were unable to displace apoA-II from the lipid/water interface. The average residue hydrophobicity of apoA-II is higher than that of apoA-I; this may contribute to the higher affinity of apoA-II for lipids compared to apoA-I. The probable helical regions in apolipoproteins A-I and A-II were located using a secondary structure prediction algorithm. The analysis suggests that the amphiphilic properties of the alpha-helical regions of apoA-I and apoA-II are probably not significantly different. Further understanding of the differences in surface activity of these apolipoproteins will require more knowledge of their secondary and tertiary structures.  相似文献   

13.
The preruminant calf (Bos spp.) is a model of considerable interest with regard to hepatic and intestinal lipoprotein metabolism (Bauchart et al., J. Lipid Res. (1989) 30, 1499–1514 and Laplaud et al., J. Lipid Res. (1990) 31, 1781–1792). As a preliminary step towards future experiments dealing with HDL metabolism in the calf, we have purified apoA-I from this animal and determined its complete amino acid sequence. Thus, approx. 10% of calf apoA-I was shown to contain a propeptide, with the sequence Arg-His-Phe-Trp-Gln-Gln. Enzymatic cleavage of apoA-I resulted in 10 proteolytic peptides. The complete apoA-I sequence was obtained after alignment of peptides on the basis of their homologies with those from rabbit apoA-I. Thus calf apoA-I consists of 241 amino acid residues, and exhibits high sequence homology with all mammalian apoA-I's studied to date. The bovine protein contained 10 hydrophobic amphipathic helical regions, occurring between residues 43–64, 65–86, 87–97, 98–119, 120–141, 142–163, 164–184, 185–206, 207–217 and 218–241. A computer-constructed phylogenetic tree showed that bovine apoA-I was more closely related to its dog counterpart, including the presence of a single methionine, than to the corresponding macaque and human proteins. Comparative predictions of the respective antigenic structures of human and bovine apoA-I's using the Hopp-Woods algorithm indicated similar positions for all 13 detectable antigenic sites, among which 7 were of identical, or closely related, amino acid composition. This finding was confirmed by demonstration of partial immunological identity between the two proteins upon immunodiffusion analysis, a result obtained using a monospecific rabbit antiserum against bovine apoA-I. Finally, comparison of sequence homology between bovine apoA-I and the lecithin : cholesterol acyl transferase (LCAT) activating region of human apoC-I suggests that several LCAT activating domains may be present in calf apoA-I.  相似文献   

14.
We produced human apolipoprotein A-I (apoA-I) in Chinese hamster ovary (CHO) cells. The CHO cells were transfected with an expression plasmid which placed the human apoA-I gene under the direction of the human metallothionein II gene promoter. Isolation of a clonal cell line resulted in high level expression of apoA-I. Greater than 30% of total protein secreted by these CHO cells was apoA-I, which enabled us to purify apoA-I with a single step purification scheme. As a result, large quantities of apoA-I can be produced and isolated without having to rely on plasma sources. Structural characterization of the recombinant apoA-I showed it to be identical to authentic apoA-I from human serum high density lipoprotein. Furthermore, we demonstrated approximately equal to 90% of the apoA-I secreted by CHO cells is processed, mature protein. A portion of the secreted recombinant apoA-I was associated with lipid and floated at a density approximately equal to 1.10 g/ml. Additional analysis identified the presence of five isoforms of apoA-I in the CHO cell conditioned medium. Processing and post-translational modification of the recombinant apoA-I occurred in the CHO cell cultures in the absence of serum components. We conclude that the human apoA-I produced by CHO cells is identical to circulating, mature apoA-I in humans and that recombinant mammalian expression offers an opportunity to investigate apoA-I processing.  相似文献   

15.
16.
In the course of an electrophoretic mutation screening program of 32,000 dried blood samples from newborns, 17 genetic variants of apolipoprotein A-I (apoA-I) were found and structurally analyzed. The following defects were identified by the combined use of high performance liquid chromatography, time-of-flight secondary ion mass spectrometry, and sequence analysis: Pro3----Arg (1 x), Pro4----Arg (1 x), Asp89----Glu (1 x), Lys107----0 (4 x), Lys107----Met (2 x), Glu139----Gly (2 x), Glu147----Val (1 x), Pro165----Arg (4 x), and Glu198----Lys (1 x). The distribution of point mutations in the apoA-I gene leading to these 9 and 11 other variants of apoA-I reported previously was statistically analyzed. Substitutions are overrepresented in the 10 amino-terminal amino acids (p less than 0.001, chi 2-test) and in residues 103-177 (p less than 0.025, chi 2-test) or residues 103-198 (p less than 0.05, chi 2-test), respectively. We further noted the following. (i) Prolines were substituted by arginine or histidine residues at a frequency much higher than expected on the basis of random nucleotide substitutions (5 out of 18 "electrically non-neutral" amino acid substitutions, p less than 0.001, chi 2-test). These substitutions are the result of transversions of cytosines contained within stretches of at least 5 consecutive cytosines in the apoA-I gene. The observed hypervariability of the apoA-I amino terminus, therefore, might be caused by a hot spot for mutation formed by the 7 subsequent cytosines in codons 3, 4, and 5. (ii) CpG dinucleotides were overrepresentatively affected by C----T transitions (5 out of 18 electrically nonneutral amino acid substitution, p less than 0.001, chi 2-test). The hypervariability of the apoA-I alpha-helical domain might therefore be caused by CpG dinucleotides predominantly occurring in codons 120-208 of apoA-I (82 out of 125). (iii) Comparison of mutation sites in the human apoA-I gene with sites of nonsynonymous substitutions revealed that amino acid substitutions found in human apoA-I were predominantly localized in areas that were little conserved during mammalian evolution. These regions may therefore represent areas of less structural constraint for the function of apoA-I.  相似文献   

17.
Repetitive DNA sequences in the bovine corticotropin-beta-lipotropin precursor gene region have been mapped and subjected to nucleotide sequence analysis. Two of the four repetitive DNA segments found are located in the 5'-flanking region, and one each within the intervening sequences. Each repetitive DNA segment contains one to three highly homologous unit sequences with an approximate length of 120 base pairs. All the unit sequences are flanked on the 3' side by tandem repeats. There are about 10(5) copies of the repetitive DNA in the bovine genome. Comparison of the bovine repetitive sequences with those of other mammalian species reveals the presence of a homologous segment of approximately 40 base pairs. This segment and the region preceding it in the bovine repetitive DNA exhibit sequence homology with the region encompassing the origin of DNA replication in papovaviruses.  相似文献   

18.
19.
Summary Actin genic regions were isolated and characterized from the heterokont-flagellated protists,Achlya bisexualis (Oomycota) andCostaria costata (Chromophyta). Restriction enzyme and cloning experiments suggested that the genes are present in a single copy and sequence determinations revealed the existence of two introns in theC. costata actin genic region. Phylogenetic analyses of actin genic regions using distance matrix and maximum parsimony methods confirmed the close evolutionary relationship ofA. bisexualis andC. costata suggested by ribosomal DNA (rDNA) sequence comparisons and reproductive cell ultrastructure. The higher fungi, green plants, and animals were seen as monophyletic groups; however, a precise order of branching for these assemblages could not be determined. Phylogenetic frameworks inferred from comparisons of rRNAs were used to assess rates of evolution in actin genic regions of diverse eukaryotes. Actin genic regions had nonuniform rates of nucleotide substitution in different lineages. Comparison of rates of actin and rDNA sequence divergence indicated that actin genic regions evolve 2.0 and 5.3 times faster in higher fungi and flowering plants, respectively, than their rDNA sequences. Conversely, animal actins evolve at approximately one-fifth the rate of their rDNA sequences.  相似文献   

20.
Nucleotide sequence of cloned cDNA of human apolipoprotein A-I.   总被引:6,自引:3,他引:3       下载免费PDF全文
P Cheung  L Chan 《Nucleic acids research》1983,11(11):3703-3715
ApoA-I is the major human HDL apoprotein. By oligonucleotide hybridization, we have isolated 5 dscDNA clones to human hepatic apo A-I mRNA. One of these clones (pA1-3) was completely sequenced. It has 878 bp plus a poly A tail of 48 and includes all the coding and 3'-untranslated regions of the mRNA and part of the 5'-untranslated region. It predicts a peptide sequence of 267 amino acids (including the 24 amino acid prepropeptides) which is very similar to the sequence reported by Brewer et al., (1978) Biochem. Biophys. Res. Commun. 80:623-630. The predicted signal peptide sequence is highly homologous to the rat apoA-I signal peptide. There is no evidence for any internally repeated segments in apoA-I either at the amino acid or at the DNA level. Using pA1-3 as a probe, we have detected on Northern gels apo A-I mRNA sequences of approximately 1100 nucleotides in human hepatic and baboon hepatic and intestinal RNAs, but not in RNAs from baboon skeletal muscle, kidney or spleen. The demonstration of apo A-I mRNA sequences in specific organs is important to our concept of "reverse cholesterol transport".  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