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.
To investigate the binding of 5′–CpG–3′ sequences by small molecules, two pyrrole (Py)–imidazole (Im) hairpin polyamides, PyImPyIm–γPyImPyIm–βDp (1) and PyIm–βIm–γPyIm–β–Im–β–Dp (2), which recognize the sequence 5′–CGCG–3′, were synthesized. The binding affinities of the 5′–CGCG–3′ sequence to the Py–Im hairpin polyamides were measured by surface plasmon resonance (SPR) analysis. SPR data revealed that dissociation equilibrium constants (Kd) of polyamides 1 and 2 were 1.1 (± 0.3) × 10–6 M and 1.7 (± 0.4) × 10–8 M, respectively. Polyamide 2 possesses great binding affinity for this sequence, 65-fold higher than polyamide 1. Moreover, when all cytosines in 5′–CpGpCpG–3′ were replaced with 5-methylcytosines (mCs), the Kd value of polyamide 2 increased to 5.8 (± 0.7) × 10–9 (M), which indicated about 3-fold higher binding than the unmethylated 5′–CGCG–3′ sequence. These results suggest that polyamide 2 would be suitable to target CpG-rich sequences in the genome.  相似文献   

2.
The quadruplex structures of the human telomere sequences AG3(T2AG3)3 I and (T2AG3)4 II were investigated in the presence of Na+ and K+ ions, through the cross-linking of adenines and guanines by the cis- and trans-[Pt(NH3)2(H2O)2](NO3)2 complexes 1 and 2. The bases involved in chelation of the cis- and trans-Pt(NH3)2 moieties were identified by chemical and 3′-exonuclease digestions of the products isolated after denaturing gel electrophoresis. These are the four adenines of each sequence and four out of the 12 guanines. Two largely different structures have been reported for I: A from NMR data in Na+ solution and B from X-ray data of a K+-containing crystal. Structure A alone agrees with our conclusions about the formation of the A1–G10, A13–G22, A1–A13 platinum chelates at the top of the quadruplex and A7–A19, G4–A19 and A7–G20 at the bottom, whether the Na+ or K+ ion is present. At variance with a recent proposal that structures A and B could be the major species in Na+ and K+ solutions, respectively, our results suggest that structure A exists predominantly in the presence of both ions. They also suggest that covalent platinum cross-linking of a human telomere sequence could be used to inhibit telomerase.  相似文献   

3.
Multiplex quantitative PCR based on novel design of fluorescent primers is described. Fluorogenic primers are labeled with a single fluorophore on a base close to the 3′ end with no quencher required. A tail of 5–7 nt is added to the 5′ end of the primer to form a blunt-end hairpin when the primer is not incorporated into a PCR product. This design provides a low initial fluorescence of the primers that increases up to 8-fold upon formation of the PCR product. The hairpin oligonucleotides (ΔG from 1.6 to –5.8 kcal/mol) may be as efficient as linear primers and provide additional specificity to the PCR by preventing primer-dimers and mispriming. Multiple fluorogenic primers were designed by specialized software and used for real-time quantitation of c-myc and IL-4 cDNAs in the presence of reference genes such as β-actin, GAPDH and 18S rRNA. Targets of 10107 copies were detected with precision in PCR using FAM-labeled primers for variable genes and JOE-labeled primers for the reference genes. This method was also used to detect single nucleotide polymorphism of the human retinal degeneration gene by allele-specific PCR with end-point detection using a fluorescent plate reader or a UV-transilluminator. We conclude that fluorogenic mono-labeled primers are an efficient and cost-effective alternative to FRET-labeled oligonucleotides.  相似文献   

4.
Experimental studies of complete mammalian genes and other genetic domains are impeded by the difficulty of introducing large DNA molecules into cells in culture. Previously we have shown that GST–Z2, a protein that contains three zinc fingers and a proline-rich multimerization domain from the polydactyl zinc finger protein RIP60 fused to glutathione S-transferase (GST), mediates DNA binding and looping in vitro. Atomic force microscopy showed that GSTZ2 is able to condense 130–150 kb bacterial artificial chromosomes (BACs) into protein–DNA complexes containing multiple DNA loops. Condensation of the DNA loops onto the Z2 protein–BAC DNA core complexes with cationic lipid resulted in particles that were readily transferred into multiple cell types in culture. Transfer of total genomic linear DNA containing amplified DHFR genes into DHFR cells by GST–Z2 resulted in a 10-fold higher transformation rate than calcium phosphate co-precipitation. Chinese hamster ovarian cells transfected with a BAC containing the human TP53 gene locus expressed p53, showing native promoter elements are active after GST–Z2-mediated gene transfer. Because DNA condensation by GST–Z2 does not require the introduction of specific recognition sequences into the DNA substrate, condensation by the Z2 domain of RIP60 may be used in conjunction with a variety of other agents to provide a flexible and efficient non-viral platform for the delivery of large genes into mammalian cells.  相似文献   

5.
Alterations of the proto-oncogene MLL (mixed lineage leukemia) are characteristic for a high proportion of acute leukemias, especially those occurring in infants. The activation of MLL is achieved either by an internal tandem duplication of 5′ MLL exons or by chromosomal translocations that create chimeric proteins with the N-terminus of MLL fused to a variety of different partner proteins. A domain of MLL with significant homology to the eukaryotic DNA methyltransferases (MT domain) has been found to be essential for the transforming potential of the oncogenic MLL derivatives. Here we demonstrate that this domain specifically recognizes DNA with unmethylated CpG sequences. In gel mobility shifts, the presence of CpG was sufficient for binding of recombinant GST–MT protein to DNA. The introduction of 5-methylCpG on one or both DNA strands precluded an efficient interaction. In surface plasmon resonance a KD of ~3.3 × 10–8 M was determined for the GSTMT/DNA complex formation. Site selection experiments and DNase I footprinting confirmed CpG as the target of the MT domain. Finally, this interaction was corroborated in vivo in reporter assays utilizing the DNA-binding properties of the MT domain in a hybrid MTVP16 transactivator construct.  相似文献   

6.
RNA polymerase II from the fission yeast Schizosaccharomyces pombe consists of 12 species of subunits, Rpb1–Rpb12. We expressed these subunits, except Rpb4, simultaneously in cultured insect cells with baculovirus expression vectors. For the isolation of subunit complexes formed in the virus-infected cells, a glutathione S-transferase (GST) sequence was fused to the rpb3 cDNA to produce GSTRpb3 fusion protein and a decahistidine-tag sequence was inserted into the rpb1 cDNA to produce Rpb1H protein. After successive affinity chromatography on glutathione and Ni2+ columns, complexes consisting of the seven subunits, Rpb1H, Rpb2, GSTRpb3, Rpb5, Rpb7, Rpb8 and Rpb11, were identified. Omission of the GST–Rpb3 expression resulted in reduced assembly of the Rpb11 into the complex. Direct interaction between Rpb3 and the other six subunits was detected by pairwise coexpression experiments. Coexpression of various combinations of a few subunits revealed that Rpb11 enhances Rpb3–Rpb8 interaction and consequently Rpb8 enhances Rpb1–Rpb3 interaction to some extent. We propose a mechanism in which the assembly of RNA poly-merase II is stabilized through multiple subunit–subunit contacts.  相似文献   

7.
Active-site-inhibited plasmin was prepared by inhibition with d-valyl-l-phenylalanyl-l-lysylchloromethane or by bovine pancreatic trypsin inhibitor (Kunitz inhibitor). Active-site-inhibited Glu-plasmin binds far more strongly to fibrin than Glu-plasminogen [native human plasminogen with N-terminal glutamic acid (residues 1–790)]. This binding is decreased by α2-plasmin inhibitor and tranexamic acid, and is, in the latter case, related to saturation of a strong lysine-binding site. In contrast, α2-plasmin inhibitor and tranexamic acid have only weak effects on the binding of Glu-plasminogen to fibrin. This demonstrates that its strong lysine-binding site is of minor importance to its binding to fibrin. Active-site-inhibited Lys-plasmin and Lys-plasminogen (Glu-plasminogen lacking the N-terminal residues Glu1–Lys76, Glu1–Arg67 or Glu1–Lys77)display binding to fibrin similar to that of active-site inhibited Glu-plasmin. In addition, α2-plasmin inhibitor or tranexamic acid similarly decrease their binding to fibrin. Glu-plasminogen and active-site-inhibited Glu-plasmin have the same gross conformation, and conversion into their respective Lys- forms produces a similar marked change in conformation [Violand, Sodetz & Castellino (1975) Arch. Biochem. Biophys. 170, 300–305]. Our results indicate that this change is not essential to the degree of binding to fibrin or to the effect of α2-plasmin inhibitor and tranexamic acid on this binding. The conversion of miniplasminogen (Glu-plasminogen lacking the N-terminal residues Glu1–Val441) into active-site-inhibited miniplasmin makes no difference to the degree of binding to fibrin, which is similarly decreased by the addition of tranexamic acid and unaffected by α2-plasmin inhibitor. Active-site-inhibited Glu-plasmin, Lys-plasmin and miniplasmin have lower fibrin-binding values in a plasma system than in a purified system. Results with miniplasmin(ogen) indicate that plasma proteins other than α2-plasmin inhibitor and histidine-rich glycoprotein decrease the binding of plasmin(ogen) to fibrin.  相似文献   

8.
The DNA repair protein O6-alkylguanine alkyltransferase (AGT) is responsible for removing promutagenic alkyl lesions from exocyclic oxygens located in the major groove of DNA, i.e. the O6 and O4 positions of guanine and thymine. The protein carries out this repair reaction by transferring the alkyl group to an active site cysteine and in doing so directly repairs the premutagenic lesion in a reaction that inactivates the protein. In order to trap a covalent AGT–DNA complex, oligodeoxyribonucleotides containing the novel nucleoside N1,O6-ethanoxanthosine (eX) have been prepared. The eX nucleoside was prepared by deamination of 3′,5′-protected O6-hydroxyethyl-2′-deoxyguanosine followed by cyclization to produce 3′,5′-protected N1,O6-ethano-2′-deoxyxanthosine, which was converted to the nucleoside phosphoramidite and used in the preparation of oligodeoxyribonucleotides. Incubation of human AGT with a DNA duplex containing eX resulted in the formation of a covalent protein–DNA complex. Formation of this complex was dependent on both active human AGT and eX and could be prevented by chemical inactivation of the AGT with O6-benzylguanine. The crosslinking of AGT to DNA using eX occurs with high yield and the resulting complex appears to be well suited for further biochemical and biophysical characterization.  相似文献   

9.
The tertiary structure of the 3′-cleaved product of the genomic hepatitis delta virus (HDV) ribozyme was solved by X-ray crystallographic analysis. In this structure, three single-stranded regions (SSrA, -B and -C) interact intricately with one another via hydrogen bonds between nucleotide bases, phosphate oxygens and 2′-OHs to form a nested double pseudoknot structure. Among these interactions, two Watson–Crick (W–C) base pairs, 726G–710C and 727G–709C, that form between SSrA and SSrC (P1.1) seem to be especially important for compact folding. To characterize the importance of these base pairs, ribozymes were subjected to in vitro selection from a pool of RNA molecules randomly substituted at positions 709, 710, 726 and 727. The results establish the importance of the two WC base pairs for activity, although some mutants are active with one G–C base pair. In addition, the kinetic parameters were analyzed in all 16 combinations with two canonical base pairs. Comparison of variant ribozymes with the wild-type ribozyme reveals that the difference in reaction rates for these variants (ΔΔG) is not simply accounted for by the differences in the stability of P1.1 (ΔΔG037). The role played by Mg2+ ions in formation of the P1.1 structure is also discussed.  相似文献   

10.
Lemontt JF  Fugit DR  Mackay VL 《Genetics》1980,94(4):899-920
The umr7–1 mutation, previously identified in a set of mutants that had been selected for defective UV-induced mutagenesis at CAN1, affects other cellular functions, including many of those regulated by the mating-type locus (MAT) in heterothallic Saccharomyces cerevisiae. The recessive umr7–1 allele, mapping approximately 20 cM distal to thr4 on chromosome III, causes clumpy growth in both a and α cells and has no apparent effect on a mating functions. However, α umr7 meiotic segregants fail to express several α-specific functions (e.g., high-frequency conjugation with a strains, secretion of the hormone α-factor and response to the hormone a-factor). In addition, α umr7 cells exhibit some a-specific characteristics, such as the barrier phenotype (Bar+) that prevents diffusion of α-factor and an increased mating frequency with α strains. The most striking property of α umr7 strains is their altered morphology, in which mitotic cells develop an asymmetric pear shape, like that of normal a cells induced to form "shmoos" by interaction with α-factor. Some a/α-specific diploid functions are also affected by umr7; instead of polar budding patterns, aumr7/umr7 diploids have medial budding like a/a, α/α and haploid strains. Moreover, aumr7/umr7 diploids have lost the ability to sporulate and are Bar+ like a or a/a strains. Revertant studies indicate that umr7–1 is a single point mutation. The umr7 mutant fails to complement mutants of both tup1 (selected for deoxythymidine monophosphate utilization) and cyc9 (selected for high iso-2-cytochrome c levels), and all three isolates have similar genetic and phenotypic properties. It is suggested that the product of this gene plays some common central role in the complex regulation of the expression of both MAT-dependent and MAT-independent functions.  相似文献   

11.
8-Oxoguanine (8-oxoG), induced by reactive oxygen species and arguably one of the most important mutagenic DNA lesions, is prone to further oxidation. Its one-electron oxidation products include potentially mutagenic guanidinohydantoin (Gh) and spiroiminodihydantoin (Sp) because of their mispairing with A or G. All three oxidized base-specific DNA glycosylases of Escherichia coli, namely endonuclease III (Nth), 8-oxoG-DNA glycosylase (MutM) and endonuclease VIII (Nei), excise Gh and Sp, when paired with C or G in DNA, although Nth is less active than the other two. MutM prefers Sp and Gh paired with C (kcat/Km of 0.24–0.26 min–1 nM–1), while Nei prefers G over C as the complementary base (kcat/Km 0.15–0.17 min–1 nM–1). However, only Nei efficiently excises these paired with A. MutY, a 8-oxoG·A(G)-specific A(G)-DNA glycosylase, is inactive with Gh(Sp)·A/G-containing duplex oligonucleotide, in spite of specific affinity. It inhibits excision of lesions by MutM from the Gh·G or Sp·G pair, but not from Gh·C and Sp·C pairs. In contrast, MutY does not significantly inhibit Nei for any Gh(Sp) base pair. These results suggest a protective function for MutY in preventing mutation as a result of A (G) incorporation opposite Gh(Sp) during DNA replication.  相似文献   

12.
Many sequence variations of the 8–17 RNA-cleaving deoxyribozyme have been isolated through in vitro selection. In an effort to understand how these sequence variations affect cleavage site selectivity, we systematically mutated the catalytic core of 8–17 and measured the cleavage activity of each mutant deoxyribozyme against all 16 possible chimeric (RNA/DNA) dinucleotide junctions. We observed sequence-function relationships that suggest how the following non-conserved positions in the catalytic core influence selectivity at the dinucleotide (5′ rN18-N1.1 3′) cleavage site: (i) positions 2.1 and 12 represent a primary determinant of the selectivity at the 3′ position (N1.1) of the cleavage site; (ii) positions 15 and 15.0 represent a primary determinant of the selectivity at the 5′ position (rN18) of the cleavage site and (iii) the sequence of the 3-bp intramolecular stem has relatively little influence on cleavage site selectivity. Furthermore, we report for the first time that 8–17 variants have the collective ability to cleave all dinucleotide junctions with rate enhancements of at least 1000-fold over background. Three optimal 8–17 variants, identified from ~75 different sequences that were examined, can collectively cleave 10 of 16 junctions with useful rates of 0.1 min−1, and exhibit an overall hierarchy of reactivity towards groups of related junctions according to the order NG > NA > NC > NT.  相似文献   

13.
We investigated 3,3′,5-tri-iodo-l-thyronine transport by human erythrocytes and by `ghosts'' prepared from these cells. Uptake of tri-iodothyronine by erythrocytes at 37°C was time-dependent with a maximum reached after 60min. Tracer analysis after incubation for 1min revealed only one saturable binding site, with Km 128±19nm (mean±s.e.m.; n=7) and Vmax. 4.6±0.7pmol of tri-iodothyronine/min per 6×107 cells. After 10min incubation Km 100±16nm (n=10) was found with Vmax. 7.7±1.2pmol of tri-iodothyronine/10min per 6×107 cells. At 0°C the uptake system is still active, with Km 132±26nm and Vmax. 1.8±0.3pmol of tri-iodothyronine/10min per 6×107 cells. The Vmax. with intact cells is 5-fold greater than the Vmax. with membranes derived from the same amount of cells when uptake studies are performed in media with similar free tri-iodothyronine concentrations. This indicates that at least 80% of tri-iodothyronine taken up by the intact erythrocytes enters the cell. This saturable uptake system can be inhibited by X-ray-contrast agents in a dose-dependent fashion. (±)-Propranolol, but not atenolol, has the same effect, indicating that the membrane-stabilizing properties of (±)-propranolol are involved. Furthermore, there is no inhibition by ouabain or vanadate, which indicates that tri-iodothyronine uptake is not dependent on the activity of Na++K+-dependent adenosine triphosphatase. We have prepared erythrocyte `ghosts'', resealed after 2.5min with 0mm-, 2mm- or 4mm-ATP inside. Inclusion of ATP and integrity of the membrane of the erythrocyte `ghosts'' were verified on the basis of an ATP-concentration-dependent functioning of the Ca2+ pump. No difference was found in the uptake of tri-iodothyronine by erythrocyte `ghosts'' with or without ATP included, indicating that uptake of tri-iodothyronine is not ATP-dependent. The following conclusions are drawn. (1) Tri-iodothyronine enters human erythrocytes. (2) There is only one saturable uptake system present for tri-iodothyronine, which is neither energy (i.e. ATP)-dependent nor influenced by the absence of an Na+ gradient across the plasma membrane. This mode of uptake of tri-iodothyronine by human erythrocytes is in sharp contrast with that of rat hepatocytes, which uptake system is energy-dependent and ouabain-sensitive [Krenning, Docter, Bernard, Visser & Hennemann (1978) FEBS Lett. 91, 113–116; Krenning, Docter, Bernard, Visser & Hennemann (1980) FEBS Lett. 119, 279–282]. (3) X-ray-contrast agents inhibit tri-iodothyronine uptake by erythrocytes in a similar fashion to that by which they inhibit the uptake of tri-iodothyronine by rat hepatocytes [Krenning, Docter, Bernard, Visser & Hennemann (1982) FEBS Lett. 140, 229–233].  相似文献   

14.
An NAD+-dependent DNA ligase from the hyperthermophilic bacterium Aquifex aeolicus was cloned, expressed in Escherichia coli and purified to homogeneity. The enzyme is most active in slightly alkaline pH conditions with either Mg2+ or Mn2+ as the metal cofactor. Ca2+ and Ni2+ mainly support formation of DNA–adenylate intermediates. The catalytic cycle is characterized by a low kcat value of 2 min–1 with concomitant accumulation of the DNAadenylate intermediate when Mg2+ is used as the metal cofactor. The ligation rates of matched substrates vary by up to 4-fold, but exhibit a general trend of T/A G/C < C/G < A/T on both the 3′- and 5′-side of the nick. Consistent with previous studies on Thermus ligases, this Aquifex ligase exhibits greater discrimination against a mismatched base pair on the 3′-side of the nick junction. The requirement of 3′ complementarity for a ligation reaction is reaffirmed by results from 1 nt insertions on either the 3′- or 5′-side of the nick. Furthermore, most of the unligatable 3′ mismatched base pairs prohibit formation of the DNAadenylate intermediate, indicating that the substrate adenylation step is also a control point for ligation fidelity. Unlike previously studied ATP ligases, gapped substrates cannot be ligated and intermediate accumulation is minimal, suggesting that complete elimination of base pair complementarity on one side of the nick affects substrate adenylation on the 5′-side of the nick junction. Relationships among metal cofactors, ligation products and intermediate, and ligation fidelity are discussed.  相似文献   

15.
The CYC7–H2 mutation causes an approximately 20-fold overproduction of iso–2–cytochromo c in a and α haploid strains of the yeast Saccharomyces cerevisiae due to an alteration in the nontranslated regulatory region that is presumably contiguous with the structural region. In this investigation, we demonstrated that heterozygosity at the mating type locus, a/α or a/a/α/α, prevents expression of the overproduction, while homozygosity, a/a and α/α, and hemizygosity, a/0 and α/0, allow full expression of the CYC7–H2 mutation, equivalent to the expression observed in a and α haploid strains. There is no decrease in the overproduction of iso-2-cytochrome c in a/α diploid strains containing either of the other two similar mutations, CYC7–H1 and CYC7–H3. It appears as if active expression of one or another of the mating-type alleles is required for the overproduction of iso-2-cytochrome c in CYC7–H2 mutants.  相似文献   

16.
The purpose of this study was to compare: a new five-layered poly (L–lactide–co–caprolactone) (PLC) membrane and small intestinal submucosa (SIS) as a control in rat urinary bladder wall regeneration. The five-layered poly (L–lactide–co–caprolactone) membrane was prepared by an electrospinning process. Adipose tissue was harvested from five 8-week old male Wistar rats. Adipose derived stem cells (ADSCs) were seeded in a density of 3×106 cells/cm2 onto PLC membrane and SIS scaffolds, and cultured for 5-7 days in the stem cell culture medium. Twenty male Wistar rats were randomly divided into five equal groups. Augmentation cystoplasty was performed in a previously created dome defect. Groups: (I) PLC+ 3×106ADSCs; (II) SIS+ 3×106ADSCs; (III) PLC; (IV) SIS; (V) control. Cystography was performed after three months. The reconstructed urinary bladders were evaluated in H&E and Masson''s trichrome staining. Regeneration of all components of the normal urinary bladder wall was observed in bladders augmented with cell-seeded SIS matrices. The urinary bladders augmented with SIS matrices without cells showed fibrosis and graft contraction. Bladder augmentation with the PLC membrane led to numerous undesirable events including: bladder wall perforation, fistula or diverticula formation, and incorporation of the reconstructed wall into the bladder lumen. The new five-layered poly (L–lactide–co–caprolactone) membrane possesses poorer potential for regenerating the urinary bladder wall compared with SIS scaffold.  相似文献   

17.
Cell types in addition to those previously described (Kruse et al. 1963. J. Nat. Cancer Inst. 31:109; Kruse and Miedema. 1965. J. Cell Biol. 27:273) were found to form multiple-layered cultures by perfusion-culture technique. Dense populations containing 43 x 106 embryonic rat muscle (NF-ER) cells, 23 x 106 diploid human tonsillar (NF-JAM) cells, 77 x 106 human pleural effusion isolate (RPMI 2650) cells, 35 x 106 embryonic diploid human lung (Flow 2000) cells, 21 x 106 bovine lung (FB4BM) cells, 108 x 106 bat lung (Tb1Lu) cells, and 81 x 106 SV-40 virus-transformed embryonic diploid human lung (WI-38VA13A) cells were obtained in 6–14 days from dilute inocula in T-60 or T-75 flasks; these were equivalent to about 4, 3, 3, 4, 2, 4, and eight monolayers, respectively. Perfusion of an NF-ER culture for 6 wk with medium plus 10% whole calf serum yielded a cell density equivalent to 12 monolayers (140 x 106 cells per T-75 flask). This culture exhibited random labeling of nuclei from bottom to top after pulsing for 90 min with thymidine-3H. Medium plus 0.1% serum maintained NF-JAM cultures at constant viable cell numbers with virtual absence of thymidine-3H labeling. Similar results were obtained with WI-38 cultures, but WI-38VA13A cells continued active DNA synthesis and mitosis in medium with 0.1% serum to form 16–20 layers of cells (191–239 x 106 cells per T-75 flask) in 27 days. WI-38VA13A cells ceased proliferation and became nonviable rapidly in serumless medium.  相似文献   

18.
Wickner RB 《Genetics》1977,87(3):441-452
Strains of Saccharomyces cerevisiae carrying a 1.4 x 106 dalton double-stranded (ds) RNA in virus-like particles (the killer plasmid or virus) secrete a toxin that is lethal to strains not carrying this plasmid (virus). The mak10 gene is one of 24 chromosomal genes (called pets, mak1, mak2,...) that are needed to maintain and replicate the killer plasmid. We report here isolation of spontaneous and induced mutants in which the killer plasmid is maintained and replicated in spite of a defect in the mak10 gene. The bypass (or suppressor) mutations in these strains are in the mitochondrial genome. Respiratory deficiency produced by various chromosomal pet mutations, by chloramphenicol, or by antimycin A, does not bypass the mak10–1 mutation. Several spontaneous mak10–1 killer strains have about 12-fold more of the killer plasmid ds RNA than do wild-type killers. Although the absence of mitochondrial DNA bypasses mak10–1, it does not bypass pets–1, mak1–1, mak3–1, mak4–1, mak5–1, mak6–1, mak7–1, or mak8–1.  相似文献   

19.
We report on the physicochemical effects resulting from incorporating a 5-(3-aminopropyl) side chain onto a 2′-deoxyuridine (dU) residue in a short DNA hairpin. A combination of spectroscopy, calorimetry, density and ultrasound techniques were used to investigate both the helix–coil transition of a set of hairpins with the following sequence: d(GCGACTTTTTGNCGC) [N = dU, deoxythymidine (dT) or 5-(3-aminopropyl)-2′-deoxyuridine (dU*)], and the interaction of each hairpin with Mg2+. All three molecules undergo two-state transitions with melting temperatures (TM) independent of strand concentration that indicates their intramolecular hairpin formation. The unfolding of each hairpin takes place with similar TM values of 64–66°C and similar thermodynamic profiles. The unfavorable unfolding free energies of 6.4–6.9 kcal/mol result from the typical compensation of unfavorable enthalpies, 36–39 kcal/mol, and favorable entropies of ~110 cal/mol. Furthermore, the stability of each hairpin increases as the salt concentration increases, the TM-dependence on salt yielded slopes of 2.3–2.9°C, which correspond to counterion releases of 0.53 (dU and dT) and 0.44 (dU*) moles of Na+ per mole of hairpin. Absolute volumetric and compressibility measurements reveal that all three hairpins have similar hydration levels. The electrostatic interaction of Mg2+ with each hairpin yielded binding affinities in the order: dU > dT > dU*, and a similar release of 2–4 electrostricted water molecules. The main result is that the incorporation of the cationic 3-aminopropyl side chain in the major groove of the hairpin stem neutralizes some local negative charges yielding a hairpin molecule with lower charge density.  相似文献   

20.
A novel replicating circular DNAzyme   总被引:6,自引:1,他引:5  
Chen F  Wang R  Li Z  Liu B  Wang X  Sun Y  Hao D  Zhang J 《Nucleic acids research》2004,32(8):2336-2341
10–23 DNAzyme has the potential to suppress gene expressions through sequence-specific mRNA cleavage. However, the dependence on exogenous delivery limits its applications. The objective of this work is to establish a replicating DNAzyme in bacteria using a single-stranded DNA vector. By cloning the 10–23 DNAzyme into the M13mp18 vector, we constructed two circular DNAzymes, C-Dz7 and C-Dz482, targeting the β-lactamase mRNA. These circular DNAzymes showed in vitro catalytic efficiencies (kcat/KM) of 7.82 × 106 and 1.36 × 107 M–1·min–1, respectively. Their dependence on divalent metal ions is similar to that found with linear 10–23 DNAzyme. Importantly, the circular DNAzymes were not only capable of replicating in bacteria but also exhibited high activities in inhibiting β-lactamase and bacterial growth. This study thus provides a novel strategy to produce replicating DNAzymes which may find widespread applications.  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