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.
Comparative 1H NMR and optical studies of the interaction between poly(dA-dT), ethidium bromide (Et) and ethidium dimer (Et2) in 0.7 M NaCl are reported as a function of the temperature. Denaturation of the complexes followed at both polynucleotide and drug levels leads to a biphasic melting process for poly(dA-dT) complexed with ethidium dimer (t1/2 = 75 degrees C; 93 degrees C) but a monophasic one in poly(dA-dT): ethidium bromide complex (t1/2 = 74 degrees C). In both cases drug signals exhibit monophasic thermal dependence (Et = 81 degrees C; Et2 = 95 degrees C). Evidence is presented showing that the ethidium dimer bisintercalates into poly(dA-dT) in high salt, based on the observation that i) dimer and monomer ring protons exhibit similar upfield shifts upon DNA binding, ii) upfield shifts of DNA sugar protons are twice as large with the dimer than with ethidium bromide. Comparison between native DNA fraction and bound drug fraction indicates that ethidium covers, n = 2.5-3 base pairs. The dimer bisintercalates and covers, n = 5.7 base pairs when the helix fraction is high but as the number of available sites decreases the binding mode changes and the drug monointercalates (n = 2.9).  相似文献   

2.
Yang XL  Robinson H  Gao YG  Wang AH 《Biochemistry》2000,39(36):10950-10957
The binding of a macrocyclic bisacridine and an antitumor intercalator ametantrone to DNA has been studied. We carried out X-ray diffraction analyses of the complexes between both intercalators and CGTACG. We have determined the crystal structure, by the multiple-wavelength anomalous diffraction (MAD) method, of bisacridine complexed with CGTA[br(5)C]G at 1.8 A resolution. The refined native crystal structure at 1.1 A resolution (space group C222, a = 29.58 A, b = 54.04 A, c = 40.22 A, and R-factor = 0.163) revealed that only one acridine of the bisacridine drug binds at the C5pG6 step of the DNA, with the other acridine plus both linkers completely disordered. Surprisingly, both terminal G.C base pairs are unraveled. The C1 nucleotide is disordered, and the G2 base is bridged to its own phosphate P2 through a hydrated Co(2+) ion. G12 is swung toward the minor groove with its base stacked over the backbone. The C7 nucleotide is flipped away from the duplex part and base paired to a 2-fold symmetry-related G6. The central four base pairs adopt the B-DNA conformation. An unusual intercalator platform is formed by bringing four complexes together (involving the 222 symmetry) such that the intercalator cavity is flanked by two sets of G x C base pairs (i.e., C5 x G8 and G6 x C7) on each side, joined together by G6 x G8 tertiary base pairing interactions. In the bisacridine-CGTACG complex, the intercalation platform is intercalated with two acridines, whereas in the ametantrone-CGTACG complex, only one ametantrone is bound. NMR titration of the bisacridine to AACGATCGTT suggests that the bisacridine prefers to bridge more than one DNA duplex by intercalating each acridine to different duplexes. The results may be relevant in understanding binding of certain intercalators to DNA structure associated with the quadruplet helix and Holliday junction.  相似文献   

3.
Iu S Babaian 《Biofizika》1991,36(1):35-38
The interaction of antitumoral drug mitoxantrone with DNA of the tumor sarcoma 45 and healthy animals liver has been investigated according to the character of changes on the absorption spectra at binding at 30 degrees C and 0.11 M NaCl. The investigation shows that the interaction of mitoxantrone with DNA of sarcoma 45 differs from that with DNA of healthy animals liver. The calculations show that the saturation stoichiometry by both DNA is one mitoxantrone molecule per 2.5 base pairs with the binding constant k = 4 x 10(5) M-1 (for binding mitoxantrone with liver DNA) and k = 3 x 10(6) M-1 (with tumor DNA). Possible reason of such a difference is discussed on the basis of structural peculiarities of tumor DNA.  相似文献   

4.
It was shown that chromatin contained in human lymphocytes has two stages of denaturation: with T(d)VII = 94.4 degrees C, Q(d)VII = 50.8 J/g DNA, and T(d)VIII = 105.1 degrees C Q(d)VIII = 44.9 J/g DNA. The peptide bioregulator prostamax causes a redistribution of heat among endotherms T(d)III and T(d)IV and a shift of both endotherms to low temperatures by 2.9 and 1.0 degrees C, respectively. It was supposed that the redistribution of heat among endotherms is connected with a partial relaxation of the 30-nm-thick fiber in the 10-nm filament. A weak decrease in T(d)VIII and T(d)VII of lymphocytes treated with prostamax compared to untreated ones is connected with small structural changes of nucleosomal organization in the 10-nm filament and 30-nm-thick fiber.  相似文献   

5.
Characterization of the genome of the basidiomycete Schizophyllum commune   总被引:8,自引:0,他引:8  
DNA of Schizophyllum commune was isolated both from mycelial cells and from protoplasts. Nuclear DNA was isolated after solubilization of the mitochondria with the detergent Nonidet. The G + C content of the nuclear DNA was 57%, calculated from its buoyant density (1.7165 g/ml) and from the Tm (77.4 degrees C in 15 mM NaCl/1.5 mM trisodium citrate). The buoyant density of the ribosomal cistrons was 1.707 g/ml. DNA isolated from purified mitochondria had a very low G + C content: 22% (rho = 1.6845 g/ml, Tm = 61.8 degrees C in 15 mM NaCl/1.5 mM trisodium citrate). Analysis of CsCl profiles and melting patterns suggested that mitochondrial DNA contains interspersed (A + T)-rich sequences. From reassociation analysis of sheared nuclear DNA the genome size of S. commune was determined to be 22.8 . 10(9) daltons. A small amount of DNA (0.5 . 10(9) daltons) bound to hydroxyapatite at zero time Cot. 7% of the genome (1.6 . 10(9) daltons) represented repetitive DNA.  相似文献   

6.
Hydroxyurea (10 mM) arrests the exponential growth of Tetrahymena by blocking DNA replication during S-phase. After removal of the hydroxyurea (HU), they have a long recovery period during which they are active in DNA synthesis. 3H-TdR uptake showed that on completion of the recovery period, the cells divide (recovery division) and enter a cell cycle which lacks G1. The frequency, size and DNA content of the extranuclear chromatin bodies (ECB) formed at this division are all markedly increased (2–4) over the corresponding values obtained from exponential growth phase controls. Microspectrophotometric analysis of macronuclear DNA content (N) coupled with the cytoplasmic dry mass (C) values suggest that specific N to C ratios (N/C) are required for the initiation of DNA replication and fission: during a normal (exponential growth) cell cycle, both N and C double, but asynchronously, so that the N/C of both post-fission-daughter cells and pre-fission cells is identical (standardized to N/C = 1) but late G1 cells have a low N/C. During a 10 hr exposure to HU, the N remains essentially the same whereas the C increases. When the HU is removed, the N increases by 4× and the C continues to increase until just prior to recovery division when it also reaches a value 4× that of the original daughter cells. Thus, the N/C = 1 is re-established. The enlarged ECB formed during recovery division may function to lower the N/C in the daughter cells, which in turn may in some way stimulate immediate DNA replication, thus eliminating G1. The elimination of G1 (and shortening in a few subsequent cell cycles) allows less time for cytoplasmic growth and results in the return of the cells to the generation time and the N and C values observed prior to the HU treatment.  相似文献   

7.
Four steroids that share the 17-hydroxy-3-oxopregna-4,6-diene structure - cyproterone acetate, chlormadinone acetate, megestrol acetate, and potassium canrenoate - have been shown previously to behave with different potency as liver-specific genotoxic agents, the response being markedly higher in female than in male rats, but similar in humans of both genders. In this study, performed to better define the relationship between chemical structure and genotoxicity, dydrogesterone (DGT) with double bonds C4=C5 and C6=C7, dienogest (DNG) with double bonds C4=C5 and C9=C10, and 1,4,6-androstatriene-17beta-ol-3-one acetate (ADT) with double bonds C1=C2, C4=C5 and C6=C7, were compared with cyproterone acetate (CPA) for their ability to induce DNA fragmentation and DNA repair synthesis in primary cultures of hepatocytes from three rats of each sex. At subtoxic concentrations, ranging from 10 to 90 microM, all four steroids consistently induced a dose-dependent increase of DNA fragmentation, which in all cases was higher in females than in males; their DNA damaging potency decreased in the order CPA > DNG > ADT > DGT. Under the same experimental conditions, the responses provided by the DNA repair-synthesis assay were positive or inconclusive in hepatocytes from female rats and consistently negative in hepatocytes from male rats. In the induction of apoptotic cells, examined in primary hepatocytes from female rats, CPA was more active than ADT and DGT, and DNG was inactive. Considered as a whole these findings suggest that a liver-specific genotoxic effect more marked in female than in male rats might be a common property of steroids with two or three double bonds.  相似文献   

8.
K P Singh  S N Raina  A K Singh 《Génome》1996,39(5):890-897
The 2C nuclear DNA amounts were determined for 99 accessions, representing 23 Arachis species from 8 of 9 taxonomic sections, and two synthetic amphidiploids. Mean 2C DNA amounts varied by 15.20%, ranging from 10.26 to 11.82 pg, between accessions of Arachis hypogaea (2n = 4x = 40). Nuclear DNA content variation (5.33-5.91 pg) was also detected among Arachis duranensis (2n = 2x = 20) accessions. The intraspecific variation in the two species may have resulted from indirect selection for favourable genome sizes in particular environmental conditions. The accessions belonging to A. hypogaea ssp. hypogaea (mean value 11.27 pg) with longer life cycle had significantly larger mean DNA content than the accessions of A. hypogaea ssp. fastigiata (mean value 10.97 pg). For 20 diploid (2n = 2x = 20) species of the genus, 2C nuclear DNA amounts ranged from approximately 3 to 7 pg. The diploid perennial species of section Arachis have about 12% more DNA than the annual species. Comparisons of DNA amounts show that evolutionary rating is not a reliable guide to DNA amounts in generic sections of the genus; lower DNA values with evolutionary advancement were found in sections Heteranthae and Triseminatae, but the same was not true for sections Arachis and Caulorrhizae. Similarly, there is evidence of significant differences in DNA content between 4 ancient sections (Procumbentes, Erectoides, Rhizomatosae, and Extranervosae) of the genus. The occurrence of genome size plasticity in both A. duranensis and A. hypogaea provides evidence that A. duranensis could be one of the diploid progenitors of A. hypogaea. The DNA content in the two synthetic amphidiploids corresponded to the sum value estimated for parental species. Key words : Arachis species, genome size, Arachis hypogaea, Arachis duranensis, intraspecific variation.  相似文献   

9.
Complementary DNAs to rat ventral prostate poly(A) RNA were cloned into pBR322 by the "dG-dC tailing" procedure. Clones containing cDNAs to the mRNAs coding for each of the three subunits of a major secretory protein (prostatein) were identified by hybrid-arrested translation. A 457-nucleotide base pair cDNA (E45) and a portion of a 365-base pair cDNA (E85) were analyzed to determine the composite complete DNA coding sequence for the Mr = 14,000 (C3) subunit of prostatein. A sequence of 12-nucleotide bases (TTTGCTGCTATG) in the signal peptide of C3 was noted to be homologous to signal peptide nucleotide sequences reported in cDNAs coding for the other two prostatein subunits, Mr = 6,000 (C1) and 10,000 (C2). Complementary DNA coding for the C3 subunit was used as a hybridization probe to screen an EcoRI rat genomic DNA library. Two unique 12-kilobase genomic clones, each containing mRNA coding sequences within 2.5-3-kilobase fragments, were identified by restriction enzyme mapping and Southern blot analysis. Restriction enzyme sites within the coding regions of both genes were analogous to the cDNA. Differences in restriction enzyme sites in regions of intervening sequences and flanking DNA established the uniqueness of the two genes. It is suggested that both genes may be transcribed in vivo.  相似文献   

10.
11.
Nuclear binding of thyroxine (T4) and triiodothyronine (T3) in mononuclear blood cells was investigated in 12 young (age 16-30 years) healthy subjects (group A), in 12 middle-aged (age 31-60 years) healthy subjects (group B) and in 12 elderly (61-90 years) healthy subjects. Serum free T3 was depressed in group C as compared to the younger age groups, whereas serum free T4 and TSH did not differ between the groups. Maximal specific nuclear binding capacity for both T4 and T3 decreased with increasing age, T4 group A: 1.2 fmol T4/100 micrograms DNA, group B: 1.2 fmol T4/100 micrograms DNA, group C: 0.7 fmol T4/100 micrograms DNA; T3 group A: 1.7 fmol T3/100 micrograms DNA, group B: 1.0 fmol T3/100 micrograms DNA, group C: 0.9 fmol T3/100 micrograms DNA. The equilibrium association constant (Ka) for T4 increased with age, group A: Ka = 3.3 X 10(9) l/mol, group B: Ka = 3.2 X 10(9) l/mol, group C: Ka = 6.4 X 10(9) l/mol, whereas Ka for nuclear binding of T3 decreased with age group A: Ka = 3.9 X 10(9) l/mol, group B: Ka = 5.9 X 10(9) l/mol, group C: Ka = 1.8 X 10(9) l/mol. We conclude that, whereas the opposite variations of nuclear capacity and binding affinity for T4 tend to preserve the nuclear T4 concentration, the nuclear T3 concentration definitely decreases with age. The unaltered serum levels of TSH suggest that the decrease of both serum levels of free T3 and the nuclear T3 concentration might represent physiologically changes in old age.  相似文献   

12.
The TaqI restriction endonuclease recognizes and cleaves the duplex DNA sequence T decreases CGA. Steady state kinetic analysis with a small oligodeoxyribonucleotide substrate showed that the enzyme obeyed Michaelis-Menten kinetics (Km = 53 nM, kcat = 1.3 min-1 at 50 degrees C and Km = 0.5 nM, kcat = 2.9 min-1 at 60 degrees C). At 0 degree C, the enzyme was completely inactive, while at 15 degrees C, turnover produced nicked substrate as the major product in excess of enzyme indicating dissociation between nicking events. Above 37 degrees C, both strands in the duplex were cleaved prior to dissociation. In contrast to the tight, temperature-dependent binding of substrate, binding of the Mg2+ cofactor was weak (Kd = 2.5 mM) and the same at either 50 degrees C or 60 degrees C. Single-turnover experiments using oligonucleotide substrate showed that hydrolysis of duplex DNA occurred via two independent nicking events, each with a first order rate constant (kst) of 5.8 min-1 at 60 degrees C and 3.5 min-1 at 50 degrees C. The pH dependence of Km (pKa = 9) and kst (pKa = 7) suggests Lys/Arg and His, respectively, as possible amino acids influencing these constants. Moreover, although kst increased significantly with pH, kcat did not, indicating that at least two steps can be rate-controlling in the reaction pathway. Binding of protein to canonical DNA in the presence of Mg2+ at 0 degree C or in the absence of Mg2+ at 50 degrees C was weak (Kd = 2.5 microM or 5,000-fold weaker than the optimal measured Km) and equal to the binding of noncanonical DNA as judged by retention on nitrocellulose. Similar results were seen in gel retardation assays. These results suggest that both Mg2+ and high temperature are required to attain the correct protein conformation to form the tight complex seen in the steady state analysis. In the accompanying paper (Zebala, J. A., Choi, J., Trainor, G. L., and Barany, F. (1992) J. Biol. Chem. 267, 8106-8116), we report how these kinetic constants are altered using substrate analogues and propose a model of functional groups involved in TaqI endonuclease recognition.  相似文献   

13.
Abstract

The interaction of the minor groove binding ligand Hoechst 33258 (Hoe) with natural DNA was investigated by high resolution titration rotational viscometry. Analysis of the concomitant DNA conformational changes was performed with two DNA samples of sufficiently different molar mass M, at 4°C, 22°C and 40°C, for Hoe/DNA-P ratios below r = 0.02. In this narrow r range several interaction modes could be resolved. The measured conformational changes were quantified in terms of relative changes of both apparent DNA persistence length, Δa/a, and hydrodynamically operative DNA contour length, ΔL/L. Δa/a(r) primarily is a measure of ligand-induced DNA helix stiffening, but both, Δa/a(r) and ΔL/L(r), generally depend also on ligand binding induced DNA bending or DNA unbending. The essential difference obviously is that Δa/a(r) is influenced by the randomly distributed helix bends and ΔL/L(r) by phased ones. The measurements performed at different temperatures deliver informations about existence and temperature dependent abolition of intrinsic helix curvature.

Both Hoe and netropsin (Nt) prefer binding to AT rich DNA segments, which are candidates for intrinsic DNA helix bends. But our data for Hoe interaction with calf thymus DNA (ctDNA) show characteristic differences to those for Nt-ctDNA interaction. Especially for Hoe, the mode of highest affinity is saturated already at a ligand concentration of roughly 1 nM (r = 0.0015 Hoe/DNA-P). It exhibits an unusually strong temperature dependence of the conformational DNA response. A Hoe-Nt competition experiment shows that Hoe binding to the sites of the very first Hoe mode is almost unaffected by bound Nt. But Hoe binding to the sites of the following Hoe modes does not occur due to the competition with Nt. Thus this mode of strongest Hoe-DNA interaction reflects a unique mechanism, possibly of high relevance for gene regulatory systems.  相似文献   

14.
The fluorescent nucleotide 2',3'-trinitrophenyl-ATP (TNP-ATP) binds at the triphosphate substrate binding site of the large (Klenow) fragment of DNA polymerase I (Pol I) as detected by direct binding studies measuring the increase in fluorescence of this ligand (n = 1.0, KD = 0.07 microM). The enzyme-TNP-ATP complex binds Mg2+ and Mn2+ tightly (KD = 0.05 microM) as measured by an increase in fluorescence on titrating with these metals. The substrate dGTP competitively displaces TNP-ATP from the enzyme (KD = 5.7 microM) de-enhancing the fluorescence. The polymerase reaction is half-maximally inhibited by 0.8 microM TNP-ATP in the presence of dATP (10 microM) as substrate. A region of the amino acid sequence of Pol I (peptide I) consisting of residues 728-777 has been synthesized and found to contain significant secondary structure by CD both in water and 50% methanol/water. In water at 3 degrees C, peptide I binds the substrate analog TNP-ATP (KD = 0.03 microM) with a stoichiometry of 0.2. In 50% methanol at 3 degrees C, peptide I binds TNP-ATP with a higher stoichiometry than in water, consistent with a 1:1 complex, but biphasically (16% of the peptide, KD = 0.09 microM; 84% of the peptide, KD = 5.0 microM), and competitively binds the Pol I substrates dATP, TTP, and dGTP (KD = 230-570 microM). Evidence from size exclusion high performance liquid chromatography suggests that these two forms of the peptide are monomer and dimer, respectively. Significantly, the peptide I-TNP-ATP complex binds duplex DNA, tightly (KD = 0.1-0.5 microM) and stoichiometrically, and single stranded DNA more weakly. The peptide I-duplex DNA complex binds both TNP-ATP (KD = 0.5-1.5 microM) and Pol I substrates (KD = 350-2100 microM) stoichiometrically. In a control experiment, a second peptide, peptide II, based on residues 840-888 of the Pol I sequence, retains secondary structure, as detected by CD, but displays no binding of TNP-ATP. The ability of peptide I, which represents only 8% of the large fragment of Pol I, to bind both substrates and duplex DNA indicates that residues 728-777 constitute a major portion of the substrate binding site of this enzyme.  相似文献   

15.
S-M Chung  J E Staub  J-F Chen 《Génome》2006,49(3):219-229
To investigate phylogenetic relationships in the genus Cucumis, 9 consensus chloroplast simple sequence repeat (ccSSR) primer pairs (ccSSR3, 9, 11, 13, 14, 17, 20, 21, and 23) were employed for DNA fragment length variation and 5 amplified fragments, ccSSR4, 12, 13, 19, and 20, were sequenced using total DNA from 13 accessions representing 7 African Cucumis species (x = 12), 3 Cucumis melo L. (x = 12) accessions, 2 Cucumis sativus L. (x = 7) accessions, and 1 Cucumis hystrix Chakr. (x = 12) accession. A Citrullus lanatus (Thunb.) Matsum. & Nakai (x = 11) accession was used as an outgroup. While fragment length analysis revealed the existence of 3 major species clusters (i.e., a group of African Cucumis species, a group composed of C. melo accessions, and a group containing C. sativus and C. hystrix species), sequence variation analysis identified 2 major species clusters (i.e., a group of African Cucumis species and a group composed of C. melo, C. sativus, and C. hystrix species). Comparative analysis using nuclear DNA (previous studies) and cpDNA sequence substitution data resulted in the placement of C. melo and C. sativus in different cluster groupings. Thus, both nuclear and cytoplasmic DNA should be employed and compared when a putative progenitor or specimens of an ancestral Cucumis species lineage is investigated. In addition, C. ficifolius (2x) and C. aculeatus (4x) of the African Cucumis species clustered together in this study. This result does not agree with reported isozyme analyses, but does agree with previously characterized chromosome homologies between these 2 species. Although African Cucumis species and C. hystrix do not share a close relationship, genetic affinities between C. sativus and C. hystrix are considerable. Combined evidence from previously published studies and data presented herein lend support to the hypothesis that C. hystrix is either a progenitor species of C. sativus or that they at least share a common ancestral lineage.  相似文献   

16.
The type II restriction endonuclease BstYI recognizes the degenerate sequence 5'-RGATCY-3' (where R = A/G and Y = C/T), which overlaps with both BamHI (GGATCC) and BglII (AGATCT), and thus raises the question of whether BstYI DNA recognition will be more BamHI-like or BglII-like. We present here the structure of BstYI bound to a cognate DNA sequence (AGATCT). We find the complex to be more BglII-like with similarities mapping to DNA conformation, domain organization, and residues involved in catalysis. However, BstYI is unique in containing an extended arm subdomain, and the mechanism of DNA capture has both BglII-like and BamHI-like elements. Further, DNA recognition is more minimal than BglII and BamHI, where only two residues mediate recognition of the entire core sequence. Taken together, the structure reveals a mechanism of degenerate DNA recognition and offers insights into the possibilities and limitations in altering specificities of closely related restriction enzymes.  相似文献   

17.
The nature of interaction of Au(III) with nucleic acids was studied by using methods such as uv and ir spectrophotometry, viscometry, pH titrations, and melting-temperature measurements. Au(III) is found to interact slowly with nucleic acids over a period of several hours. The uv spectra of native calf-thymus DNA 9pH 5.6 acetate buffer containing (0.01M NaCIO4) showed a shift in λ max to high wavelengths and an increase in optical density at 260 nm. There was a fourfold decrease in viscosity (expressed as ηsp/c). The reaction was faster at pH 4.0 and also with denatured DNA (pH 5.6) and whole yeast RNA (pH 5.6). The order of preference of Au(III) (as deduced from the time of completion of reaction) for the nucleic acids in RNA > denatured DNA > DNA. The reaction was found to be completely reversible with respect KCN. Infrared spectra of DNA-Au(III) complexes showed binding to both the phosphate and bases of DNA. The same conclusions were also arrived at by melting-temperature studies of Au(III)-DNA system. pH titrations showed liberation of two hydroxylions at r = 0.12 [r = moles of HAuCl4 added per mole of DNA-(P)] and one hydrogen ion at r = 0.5. The probable binding sites could be N(1)/N(7) of adenine, N(7) and/or C(6)O of guanine, N(3) of cytosine and N(3) of thymine. DNAs differing in their (G = C)-contents [Clostridium perfingens DNA(G = C, 29%), salmon sperm DNA (G + C, 42%) and Micrococcus lysodeikticus DNA(G + C, 29%), salmon sperm DNA (G = C, 72%)] behaved differently toward Au(III). The hyperchromicity observed for DNAs differing in (G + C)-content and cyanide reversal titrations indicate selectivity toward ( A + T)-rich DNA at lw values of r. Chemical analysis and job's continuous variation studies indicated the existence of possible complexes above and below r = 1. The results indicate that Au(III) ions probably bind to hte phosphate group in the initial stages of the reaction, particularly at low values of r, and participation of the base interaction also increases. Cross-linking of the two strands by Au(III) may take place, but a complete collapse of the doulbe helix is not envisaged. It is probable that tilting of the bases or rotaiton of the bases around the glucosidic bond, resulting in a significant distrotion of the double helix, might take place due to binding of Au(III) to DNA.  相似文献   

18.
Nuclear 1C DNA content in haploid megagametophyte tissue of 18 North American and one exotic Pinus species was determined using scanning microspectrophotometry. The nuclear DNA content in root meristematic cells of Zea mays L. ssp. mays, inbred line Va35 (4C = 10.31 pg) was used as a standard. DNA content measured by microspectrophotometry was verified using laser flow cytometry with two additional standards, Hordeum vulgare cv. Sultan (2C = 11.12 pg) and P. eldarica (2C = 47.30 pg). DNA values obtained by both methods were significantly correlated (r = 0.987). The 1C nuclear DNA content ranged from 21 pg to 31 pg. The ratio of DNA content in embryo tissue of P. eldarica to that in megagametophyte tissue was 1.72 by scanning microspectrophotometry and 1.74 by laser flow cytometry. To date, this is the most comprehensive data set available for North American Pinus species. Relationships between genome size of 18 North American Pinus species and climatic factors and indices of growth were investigated using regression and correlation analyses. Positive correlations were observed between nuclear DNA content and growth indices, minimum seed-bearing age, and seed dimensions. Strong negative correlations were observed between nuclear DNA content and two climatic factors, the lowest mean annual and monthly precipitation (excluding January) and the highest mean monthly spring air temperature. These correlations suggest that the large genome size and its variation in Pinus are adapted responses to the habitats of these species.  相似文献   

19.
20.
In this study, chromosome number and ploidy levels of Ipheion uniflorum cv. "Wisley Blue" (spring starflower) were determined. In meristematic root tip cells, chromosome number was found as 2n = 12 and 4n = 24. The ratios of diploid and tetraploid cells were found as 80.74% and 19.26%, respectively. In differentiated root tissues and mature leaf tissues ploidy levels were analysed by flow cytometry and polysomaty were found in both organs. In differentiated root tissues, ploidy levels were found as 2C, 4C, 8C and 16C DNA. In root tissues percentages of 2C, 4C, 8C and 16C nuclear DNA content were observed as 57.2%, 33.1%, 2.47% and 7.23%, respectively. In mature leaf tissues, ploidy levels were determined 2C, 4C, 8C and 16C DNA. In this tissue the frequency of 4C DNA was found very higher (74.3%) and 2C DNA content was determined as 19.2%. In mature leaf tissue, 8C and 16C nuclear DNA contents were observed as 2.72% and 3.78%, respectively. When nuclear DNA contents in leaves and roots were compared, an apparent difference in 2C and 4C DNA contents was found.  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