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.
The complexity of nuclear RNA, poly(A)hnRNA, poly(A)mRNA, and total poly(A)RNA from mouse brain has been measured by saturation hybridization with nonrepeated DNA. These DNA populations were complementary, respectively, to 21, 13.5, 3.8, and 13.3% of the DNA. From the RNA Cot required to achieve half-sturation, it was estimated that about 2.5–3% of the mass of total nuclear RNA constituted most of the complexity. Similarly, complexity driver molecules constituted 6–7% of the mass of the poly(A)hnRNA. 75–80% of the poly(A)mRNA diversity is contained in an estimated 4–5% of the mass of this mRNA. Poly(A)hnRNA constituted about 20% of the mass of nuclear RNA and was comprised of molecules which sedimented in DMSO-sucrose gradients largely between 16S and 60S. The number average size of poly(A)hnRNA determined by sedimentation, electron microscopy, or poly(A) content was 4200–4800 nucleotides. Poly(A)mRNA constituted about 2% of the total polysomal RNA, and the number average size was 1100–1400 nucleotides. The complexity of whole cell poly(A)RNA, which contains both poly(A)hnRNA and poly(A)mRNA populations, was the same as poly(A)hnRNA. This implies that cytoplasmic polyadenylation does not occur to any apparent qualitative extent and that poly(A)mRNA is a subset of the poly(A)hnRNA population. The complexity of poly(A)hnRNA and poly(A)mRNA in kilobases was 5 × 105 and 1.4 × 105, respectively. DNA which hybridized with poly(A)mRNA renatures in the presence of excess total DNA at the same rate as nonrepetitive tracer DNA. Hence saturation values are due to hybridization with nonrepeated DNA and are therefore a direct measure of the sequence complexity of poly(A)mRNA. These results indicate that the nonrepeated sequence complexity of the poly(A)mRNA population is equal to about one fourth that observed for poly(A)hnRNA.  相似文献   

2.
The binding of a dimeric distamycin analog (Pt–bis–Dst) to poly[d(A–T)]poly[d(A–T)], poly(dA)poly(dT), and duplex O23 with the sequence 5’-GCCAATATATATATATTATTAGG-3’, which occurs at the origin of replication (OriS) of the herpes simplex virus, was studied via UV and CD spectroscopy. The synthetic polyamide differs from the natural antibiotic in having two distamycin moieties that are linked via a glycine cis-diamino platinum group. The Pt–bis–Dst binding to poly[d(A–T)]poly[d(A–T)] and poly(dA)poly(dT) reached saturation at approximately one ligand molecule per eight bp. As the ligand–base pair ratio further increased, the maximum wavelength band tended to shift toward longer wavelengths in the CD spectra of complexes with poly[d(A–T)]poly[d(A–T)] and a shoulder appeared in the 290–310 nm spectral region that was absent from the CD spectra of complexes with lower ligand coverages. At higher ligand–oligonucleotide molar ratios, Pt–bis–Dst could bind to poly[d(A–T)]poly[d(A–T)] in the form of hairpins or associations that result from interactions between the distamycin moieties of two neighbor Pt–bis–Dst molecules. The structures of the complexes were stabilized by interactions between the pirrolcarboxamide moieties of two Pt–bis–Dst molecules absorbed on adjacent overlapping binding sites. The interactions could also be responsible for the concentration-dependent spectral changes that were observed during the formation of a complex between Pt–bis–Dst and poly[d(A–T)]poly[d(A–T)]. Spectral changes were almost absent in the case of Pt–bis–Dst binding to poly(dA)poly(dT). The binding of Pt–bis–Dst to duplex O23 reached saturation at two ligand molecules per duplex, which contained a cluster of 18 AT pairs. At higher molar-concentration ratios, duplex CD spectra underwent changes similar to those that were observed for Pt–bis–Dst binding to poly[d(A–T)]poly[d(A–T)]. Testing Pt–bis–Dst for antiviral activity identified 1.5 μg/mL as a concentration that halved the cytopathic effect of the herpes simplex virus on Vero E6 cells; the selectivity index of antiviral action was 65; cytotoxicity was relatively low. The Pt–bis–Dst concentration that caused the death of approximately half of the cells was estimated at 100 μg/mL.  相似文献   

3.
The relative abundance of the mRNAs encoding the membrane (mu m) and secreted (mu s) forms of immunoglobulin mu heavy chain is regulated during B-cell maturation by a change in the mode of RNA processing. Current models to explain this regulation involve either competition between cleavage-polyadenylation at the proximal (mu s) poly(A) site and cleavage-polyadenylation at the distal (mu m) poly(A) site [poly(A) site model] or competition between cleavage-polyadenylation at the mu s poly(A) site and splicing of the C mu 4 and M1 exons, which eliminates the mu s site (mu s site-splice model). To test certain predictions of these models and to determine whether there is a unique structural feature of the mu s poly(A) site that is essential for regulation, we constructed modified mu genes in which the mu s or mu m poly(A) site was replaced by other poly(A) sites and then studied the transient expression of these genes in cells representative of both early- and late-stage lymphocytes. Substitutions at the mu s site dramatically altered the relative usage of this site and caused corresponding reciprocal changes in the usage of the mu m site. Despite these changes, use of the proximal site was still usually higher in plasmacytomas than in pre-B cells, indicating that regulation does not depend on a unique feature of the mu s poly(A) site. Replacement of the distal (mu m) site had no detectable effect on the usage of the mu s site in either plasmacytomas or pre-B cells. These findings are inconsistent with the poly(A) site model. In addition, we noted that in a wide variety of organisms, the sequence at the 5' splice junction of the C mu 4-to-M1 intron is significantly different from the consensus 5' splice junction sequence and is therefore suboptimal with respect to its complementary base pairing with U1 small nuclear RNA. When we mutated this suboptimal sequence into the consensus sequence, the mu mRNA production in plasmacytoma cells was shifted from predominantly mu s to exclusively mu m. This result unequivocally demonstrated that splicing of the C mu 4-to-M1 exon is in competition with usage of the mu s poly(A) site. A key feature of this regulatory phenomenon appears to be the appropriately balanced efficiencies of these two processing reactions. Consistent with predictions of the mu s site-splice model, B cells were found to contain mu m precursor RNA that had undergone the C mu 4-to-M1 splice but had not yet been polyadenylated at the mu m site.  相似文献   

4.
The poly(A) segment increases in length as the complexity of the organism increases. The shortest poly(A) size (10–15 AMP units) exist on mRNA from bacteria. The largest poly(A) segments (180–200 AMP units) exist in highly differentiated tissue. The natural log of average poly(A) size has evolved in a linear manner with the time of evolution of different organisms. This relationship places the origin of mRNA, and possibly life, at 3·85 ± 0.2 billion years ago. The poly(A) length of mitochondrial mRNA places the origin of mitochondria within eucaryotes at 2· 1 billion years ago.  相似文献   

5.
6.
Using the presence of poly(A) tracts as a marker for mRNA, we have examined the distribution of this class of RNA between polysomes and free RNP particles. This has been done in mature oocytes and in embryos aged for various times from fertilization through to hatching of a larva. The proportion of ribosomes that are in polysomes to those that are not has been calculated. In mature oocytes, 58% of the poly(A)+ RNA and 72% of the ribosomes are not in polysomes. By 1 hr, this drops to 51% of the poly(A)+ RNA and 48% of the ribosomes. By 7 hr, a plateau is reached: 30% of each are not in polysomes. The poly(A)+ RNA in the cytoplasm of oocytes and 1-hr embryos is found in particles with an average size of 50S and a range of 30–70S. The poly(A)+ RNA ranges in size from 7 to 40S, with an average size of 22S. The polyA from this RNA is 50–200 nucleotides long with an average of 115 nucleotides. These data have allowed us to calculate that 1–2% of the total RNA is poly(A)+ RNA.  相似文献   

7.
Abstract

Fourth derivative spectrophotometry has been applied to monitor conformational isomerizations of polynucleotides for the first time. The transitions studied have been the B-A and A-X isomerizations of poly(dA-dT) and the B-X one of poly(amino2dA-dT). Parameters obtained from the fourth derivative spectra have been used to follow these conformational changes. The A form of poly(dA-dT) has been characterized by a new fourth derivative peak at 293.0 nm which can be associated to interstrand adenine-adenine interactions. Furthermore, some of the fourth derivative peaks in the long wavelength region (270–310 nm) can be related to stacking interactions present in the polynucleotide double helices. The tentative assignment of these peaks, particularly that at 299.0 nm in the derivative spectra of poly(amino2dA-dT), to n→π* electronic transitions is discussed.  相似文献   

8.
The interaction of purine with DNA, tRNA, poly A, poly C, and poly A. poly U complex was investigated. In the presence of purine, the nucleic acids in coil form (such as denatured DNA, poly A and poly C in neutral solutions, or tRNA) have lower optical rotations. In addition, hydrodynamic studies indicate that in purine solutions the denatured DNA has a higher viscosity and a decreased sedimentation coefficient. These findings indicate that through interaction with purine, the bases along the poly-nucleotide chain are unstacked and are separated farther from each other, resulting in increased assymmetry (and possibly volume) of the whole polymer. Thus, the de-naturation effect of purine reported previously can be explained by this preferential interaction of purine with the bases of nucleic acids in coil form through a hydrophobic-costacking mechanism. Results from studies on optical rotation and helix-coil transition show that the interaction of purine is greater with poly A than with poly C. The influence of temperature, Mg++ concentration, ionic strength, and purine concentration on the effect of purine on nucleic acid conformation has also been investigated. In all these situations the unraveling of nucleic acid conformation occurs at much lower temperatures (20–40°C lower) in the presence of purine (0.2–0.6M).  相似文献   

9.
During collective cell migration, the intercellular forces will significantly affect the collective migratory behaviors. However, the measurement of mechanical stresses exerted at cell–cell junctions is very challenging. A recent experimental observation indicated that the intercellular adhesion sites within a migrating monolayer are subjected to both normal stress exerted perpendicular to cell–cell junction surface and shear stress exerted tangent to cell–cell junction surface. In this study, an interfacial interaction model was proposed to model the intercellular interactions for the first time. The intercellular interaction model-based study of collective epithelial migration revealed that the direction of cell migration velocity has better alignment with the orientation of local principal stress at higher maximum shear stress locations in an epithelial monolayer sheet. Parametric study of the effects of adhesion strength indicated that normal adhesion strength at the cell–cell junction surface has dominated effect on local alignment between the direction of cell velocity vector and the principal stress orientation, while the shear adhesion strength has little effect, which provides compelling evidence to help explain the force transmission via cell–cell junctions between adjacent cells in collective cell motion and provides new insights into “adhesive belt” effects at cell–cell junction.  相似文献   

10.
T E Gunter  K K Gunter 《Biopolymers》1972,11(3):667-678
Thermal denaturation of DNA's and the corresponding helix–coil transformation of artificial polyribonucleic and polydeoxyribonucleic acids have been studied extensively both theoretically1–13 and experimentally. 14–30 Much less work has been carried out on the properties of these polynucleic acids at high pressure, and in particular, on the presure dependence of the helix–coil transition temperature.31–33 Light-scattering techniques have been used in this study to measure the pressure dependence of the helix–coil transition temperature of the two- and three-stranded helices of polyriboadenylic and polyribouridilic acids and of calf thymus DNA. From the slopes of the transition temperature vs. pressure curves and heats of transition obtained from the literature,20,34 the following volume changes from these helix–coil transitions have been obtained: (a) ?0.96 cc/mole of nucleotide base pairs for the poly (A + U) transition, (b) +0.35 cc/mole of nucleotide base trios for the poly (A + 2U) transition, and (c) +2.7 cc/mole of nucleotide base pairs for the DNA transition. The relative magnitudes and signs of these volume changes which show that poly (A + U) is destabilized by increased pressure, whereas poly (A + 2U) and calf thymus DNA are stabilized by increased pressure, indicates that further development of the helix–coil transition theory for polynucleotides is needed.  相似文献   

11.
Extrinsic Cotton effects of proflavine bound to polynucleotides   总被引:1,自引:0,他引:1  
A Blake  A R Peacocke 《Biopolymers》1967,5(4):383-397
The magnitude of the Cotton effect of proflavine which is bound to RNA or to denatured DNA depends on the ratio of bound proflavine to nucleic acid base. A statistical treatment which explains this behavior has been fitted to the experimental curves and indicates that optical activity arises through interaction between two or more bound proflavine molecules. The corresponding requirement with double helical DNA is for interaction between 3–4 proflavine molecules. Although proflavine binds to denatured DNA at pH 2.8, as shown by the shift of the proflavine spectrum, the strong binding process is absent, and to this is attributed the absence of the Cotton effect at low pH. Studies on the Cotton effects of proflavine bound to poly A and poly U at neutral pH, to poly A at acid pH and to poly (A + U) allow the generalization that a relatively rigid configuration of the binding macromolecule is required for the induction of these extrinsic Cotton effects.  相似文献   

12.
13.
Highly efficient organic–inorganic hybrid solar cells of Si‐poly(3,4‐ethylenedioxythiophene):polystyrene sulfonate (PEDOT:PSS) have been demonstrated by simultaneous structural, electrical, and interfacial engineering with low processing temperature. Si substrate has been sculpted into hierarchical structure to reduce light reflection loss and increase interfacial junction area at the same time. Regarding the electrical optimization, highly conductive organic PEDOT:PSS layer has been formulated with low sheet resistance. It is argued that the sheet resistance, rather than conductivity, is the primary parameter for the high efficiency hybrid cells, which leads to the optimization of thickness, i.e., thick enough to have low sheet resistance but transparent enough to pass the incident sunlight. Finally, siloxane oligomers have been inserted into top/bottom interfaces by contact‐printing at room ambient, which suppresses carrier recombination at interfaces and reduces contact resistance at bottom electrode. Contrary to high‐temperature doping (for the formation of front surface or back surface fields), wet solution processes or vacuum‐based deposition, the contact‐printing can be done at room ambient to reduce carrier recombination at the interfaces. The high efficiency obtained with low processing temperature can make this type of cells be a possible candidate for post‐Si photovoltaics.  相似文献   

14.
H Slegers  W Fiers 《Biopolymers》1973,12(9):2007-2021
MS2 RNA, which sediments at 27S in a neutral buffer, can be converted to a compact 57S conformation at pH 3.8. Requirements for this conversion, besides protonation, are small concentrations of Mg++ ions and a low ionic strength. On the other hand, after heating in the presence of EDTA and at low ionic strength, the RNA can be unfolded to an 11.7S form at pH 6.8 and to 10.5S at pH 3.8. The compact 57S form has lost at least 50% of its secondary structure, as determined by its hypochromicity. It corresponds to a monomer species, as will be shown in a following paper (XXIV). Comparative studies with the homopolymers poly A and poly C and with the heteropolymers poly A,U, poly A,C, and poly A,G indicate that the interactions involved in the acid RNA conformation are not simply explainable by the known interactions of the A–A+, C–C+, and/or A–C+ type.  相似文献   

15.
The human albumin gene spans 16,961 nucleotides from the putative "Cap" site to the first poly(A) addition site. It is split into 15 exons by 14 intervening sequences which are symmetrically placed within the three domains of albumin. The 5' region is highly conserved up to position -250 and contains the putative TATA (-32) and CAT (-88) boxes. A consensus 5' splice sequence reads /GTAGAGT while the 3' splice sequence is pyrimidine rich and contains CTAG/ at the splice junction. The gene contains three polyadenylation signals, and this 3' region presumably arose by triplication of a shorter fragment prior to mammalian radiation. The albumin gene exhibits a high degree of DNA polymorphism and appears to have been recently invaded by Alu repetitive sequences.  相似文献   

16.
17.
The enthalpy change accompanying the double helix–coil transition of polyriboadenylic acid (poly A) in aqueous solution has been measured optically and calorimetrically in the pH range 5.7–4.5. The course of this cooperative transition was followed optically by measuring changes in ultraviolet absorption as a function of temperature at different pH values, and calorimetrically by determining the heat capacity of the solution through the transition region. From the latter measurements, the enthalpy of transition was calculated. It is shown, that ΔH is dependent on pH as it is expected from the influence of protonation of the double helix of poly A.  相似文献   

18.
19.
The single crystal structure of the methylated sequence d(CCGGTACm(5)CGG) has been solved as an antiparallel stacked X Holliday junction to 1.5 A resolution. When compared with the parent nonmethylated d(CCGGTACCGG) structure, the duplexes are translated by 3.4 A along the helix axis and rotated by 10.8 degrees relative to each other, rendering the major grooves more accessible overall. A Ca(2+) complex is seen in the minor groove opposite the junction but is related to the B conformation of the stacked arms. At the junction itself, the hydrogen bond from the N4 nitrogen of cytosine C8 to the C7 phosphate at the crossover in the parent structure has been replaced by a water bridge. Thus, this direct interaction is not absolutely required to stabilize the junction at the previously defined ACC trinucleotide core. The more compact methylated junction forces the Na(+) of the protected central cavity of the nonmethylated junction into a solvent cluster that spans the space between the junction crossover and the stacked arms. A series of void volumes within the methylated and the nonmethylated structures suggests that small monovalent cations can fill and vacate this central cavity without the need to unfold the four-stranded Holliday junction completely.  相似文献   

20.
Psi compaction of poly[d(AT)].poly[d(AT)]   总被引:1,自引:0,他引:1  
Y A Shin  S L Feroli  G L Eichhorn 《Biopolymers》1986,25(11):2133-2148
The compaction of poly[d(A–T)] · poly[d(A–T)] by Co(III) is accompanied by the formation of ψ(+)- and ψ(-)-structures. The chirality of the ψ-structure depends on the Co(III) concentration, ionic strength, temperature, pH, and the chain length of the polymer. The two forms can be readily interconverted by manipulating these factors. Phase diagrams have been constructed that demonstrate the regions of stability of the enantiomers as a function of two variables, while other factors are held constant. At critical points in the phase diagram the two forms are in such unstable equilibrium that mechanical motion will cause ψ(+) ? ψ(-) interconversion. The formation of both ψ(+)- and ψ(-)-structures by the action of Co(III) on poly[d(A–T)] · poly[d(A–T)] contrasts markedly with the behavior of poly[d(G–C)] · poly[d(G–C)] in similar circumstances by forming only the ψ(+)-structure and that of native DNA to produce no ψ at all. Thus the base sequence is important in determining the structure of chirally associated DNA molecules.  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