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 10 毫秒
1.
2.
3.

Background

Although the rat is extensively used as a laboratory model, the inability to utilize germ line-competent rat embryonic stem (ES) cells has been a major drawback for studies that aim to elucidate gene functions. Recently, zinc-finger nucleases (ZFNs) were successfully used to create genome-specific double-stranded breaks and thereby induce targeted gene mutations in a wide variety of organisms including plants, drosophila, zebrafish, etc.

Methodology/Principal Findings

We report here on ZFN-induced gene targeting of the rat interleukin 2 receptor gamma (Il2rg) locus, where orthologous human and mouse mutations cause X-linked severe combined immune deficiency (X-SCID). Co-injection of mRNAs encoding custom-designed ZFNs into the pronucleus of fertilized oocytes yielded genetically modified offspring at rates greater than 20%, which possessed a wide variety of deletion/insertion mutations. ZFN-modified founders faithfully transmitted their genetic changes to the next generation along with the severe combined immune deficiency phenotype.

Conclusions and Significance

The efficient and rapid generation of gene knockout rats shows that using ZFN technology is a new strategy for creating gene-targeted rat models of human diseases. In addition, the X-SCID rats that were established in this study will be valuable in vivo tools for evaluating drug treatment or gene therapy as well as model systems for examining the treatment of xenotransplanted malignancies.  相似文献   

4.
The association constants (K) of nucleic acid monomers with a series of water-soluble bis-porphyrins (bisMC1, bisMC3, bisMC5, bisMC7, and bisMC11) in which two porphyrin units were linked by a methylene chain of various lengths were estimated spectrophotometrically. Among the bis-porphyrins, the K values are similar for each nucleic acid monomer, indicating that the bridging chain length does not affect the association of the bis-porphyrins with the nucleic acid monomers. The melting curves of poly(dA)-poly(dT) in the presence of bisMC3 or bisMC5 were found to be biphasic, suggesting that bisMC3 and bisMC5 are bound to poly(dA)-poly(dT) with a binding mode different from the groove binding exhibited by the corresponding porphyrin monomers. A negative-induced CD peak in the Soret region of bisMC3 and bisMC5 with poly(dA)-poly(dT) is observed and the visible spectral changes of bisMC3 and bisMC5 upon addition of poly(dA)-poly(dT) are accompanied by a large red shift of the Soret band (bisMC3: 21 nm, bisMC5: 23 nm) with substantial hypochromicity (bisMC3: 49%, bisMC5: 40%). Therefore, it is reasonable to conclude that both of the porphyrin units of bisMC3 and bisMC5 intercalate into poly(dA)-poly(dT). In contrast to poly(dA)-poly(dT), the melting curves of poly(dA·dT)2 in the presence of the bis-porphyrins did not show such biphasic behavior. Together with the CD and visible absorption data, it is certain that these bis-porphyrins do not intercalate into poly(dA·dT)2.  相似文献   

5.
Coralyne is a small crescent-shaped molecule known to intercalate duplex and triplex DNA. We report that coralyne can cause the complete and irreversible disproportionation of duplex poly(dT)·poly(dA). That is, coralyne causes the strands of duplex poly(dT)·poly(dA) to repartition into equal molar equivalents of triplex poly(dT)·poly(dA)·poly(dT) and poly(dA). Poly(dT)·poly(dA) will remain as a duplex for months after the addition of coralyne, if the sample is maintained at 4°C. However, disproportionation readily occurs upon heating above 35°C and is not reversed by subsequent cooling. A titration of poly(dT)·poly(dA) with coralyne reveals that disproportionation is favored by as little as one molar equivalent of coralyne per eight base pairs of initial duplex. We have also found that poly(dA) forms a self-structure in the presence of coralyne with a melting temperature of 47°C, for the conditions of our study. This poly(dA) self-structure binds coralyne with an affinity that is comparable with that of triplex poly(dT)·poly(dA)·poly(dT). A Job plot analysis reveals that the maximum level of poly(dA) self-structure intercalation is 0.25 coralyne molecules per adenine base. This conforms to the nearest neighbor exclusion principle for a poly(dA) duplex structure with A·A base pairs. We propose that duplex disproportionation by coralyne is promoted by both the triplex and the poly(dA) self-structure having binding constants for coralyne that are greater than that of duplex poly(dT)·poly(dA).  相似文献   

6.
7.
The effect of hydrostatic pressure upon the DNA duplex, poly(dA)poly(dT), and its component single strands, poly(dA) and poly(dT) has been studied by fourier-transform infrared spectroscopy (FT-IR). The spectral data indicate that at 28 degrees C and pressures up to 12 kbar (1200 MPa) all three polymers retain the B conformation. Pressure causes the band at 967 cm(-1), arising from water-deoxyribose interactions, to shift to higher frequencies, a result consistent with increased hydration at elevated pressures. A larger pressure-induced frequency shift in this band is observed in the single stranded polymers than in the double stranded molecule, suggesting that the effect of pressure on the hydration of single strands may be greater than upon a double stranded complex. A pressure-dependent hypochromicity in the bands attributed to base stacking indicates that pressure facilitates the base stacking in the three polymers, in agreement with previous assessments of the importance of stacking in the stabilization of DNA secondary structure at ambient and high pressures.  相似文献   

8.
The effect of berenil on plasmid DNA replication was studied on pBR322-derived plasmids containing poly(dA)poly(dT) sequences. In comparison to the parental plasmid pBR322, plasmid pKH47 harboring 100 bp of poly(dA)poly(dT) at the PvuII site showed a decrease in plasmid yield in the presence of berenil. This effect was also observed in pVL26, a related plasmid in which the location of the poly(dA)poly(dT) region had been shifted to the EcoRV site in pBR322. [(3)H]Thymidine incorporation experiments indicated that DNA synthesis may be affected in these plasmids in the presence of the drug. Bromodeoxyuridine incorporation experiments coupled to Cs(2)SO(4) equilibrium density gradient centrifugation indicated that the lower plasmid yield was due to an inhibition of DNA replication by berenil. We have also found that berenil induces DNA degradation in plasmids containing the homopolymer. Our studies strongly suggest that the effect of berenil on plasmid replication and DNA stability results from its binding to the poly(dA)poly(dT) region present in these plasmids. Moreover, we have found a correlation between the position of the poly(dA)poly(dT) region and this inhibitory effect. Thus, plasmid pKH47, containing the poly(dA)poly(dT) region most proximal to the origin of pBR322 replication, was most severely affected.  相似文献   

9.
The binding of propidium to poly(dA).poly(dT) [poly(dA.dT)] and to poly[d(A-T)].poly[d(A-T)] [poly[d(A-T)2]] has been compared under a variety of solution conditions by viscometric titrations, binding studies, and kinetic experiments. The binding of propidium to poly[d(A-T)2] is quite similar to its binding to calf thymus deoxyribonucleic acid (DNA). The interaction with poly(dA.dT), however, is quite unusual. The viscosity of a poly(dA.dT) solution first decreases and then increases in a titration with propidium at 18 degrees C. The viscosity of poly[d(A-T)2] shows no decrease in a similar titration. Scatchard plots for the interaction of propidium with poly(dA.dT) show the classical upward curvature for positive cooperativity. The curvature decreases as the temperature is increased in binding experiments. A van't Hoff plot of the observed binding constants yields an apparent positive enthalpy of approximately +6 kcal/mol for the propidium-poly(dA.dT) interaction. Propidium binding to poly[d(A-T)2] shows no evidence for positive cooperativity, and the enthalpy change for the reaction is approximately -9 kcal/mol. Both the magnitude of the dissociation constants and the effects of ionic strength are quite similar for the dissociation of propidium from poly(dA-T)2] and from poly[d(A-T)2], suggesting that the intercalated states are similar for the two complexes. The observed association reactions, under pseudo-first-order conditions, are quite different. Plots of the observed pseudo-first-order association rate constant vs. polymer concentration have much larger slopes for propidium binding to poly[d(A-T)2] than to poly(dA.dT).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
11.
R Losa  S Omari    F Thoma 《Nucleic acids research》1990,18(12):3495-3502
It was suggested that poly(dA).poly(dT) rich sequences in yeast Saccharomyces cerevisiae act as elements of constitutive promoters by exclusion of nucleosomes (Struhl, K. (1985). Proc. Natl. Acad. Sci. USA 82, 8419-8423). We have mapped the chromatin structure of the pet56-his3-ded1 region in minichromosomes and show that the poly(dA).poly(dT) sequences are located in nuclease sensitive regions. DNA fragments from the nuclease sensitive promoter region of DED1 were used for nucleosome reconstitution in vitro. We show that all sequences can form nucleosome cores and that the poly(dA).poly(dT) sequence can be incorporated in nucleosome cores. The results suggest that the nuclease sensitivity found in vivo is not established by poly(dA).poly(dT) mediated exclusion of nucleosomes.  相似文献   

12.
Isotherms of the EtBr adsorption on native and denatured poly(dA)poly(dT) in the temperature interval 20–70°C were obtained. The EtBr binding constants and the number of binding sites were determined. The thermodynamic parameters of the EtBr intercalation complex upon changes of solution temperature 20–48°C were calculated: 1.0·106 M−1K≤1.4·106 M−1, free energy ΔG o=−8.7±0.3 kcal/mol, enthalpy ΔH o≅0, and entropy ΔS o=28±0.5 cal/(mol deg). UV melting has shown that the melting temperature (T m) of EtBr-poly(dA)poly(dT) complexes (μ=0.022,4.16·10−5 M EtBr) increased by 17°C as compared with the ΔT m of free homopolymer, whereas the half-width of the transition (T m) is not changed. It was shown for the first time that EtBr forms complexes of two types on single-stranded regions of poly(dA)poly(dT) denatured at 70°C: strong (K 1=1.7·105 M−1; ΔG o=−8.10±0.03 kcal/mol) and weak (K 2=2.9·103 M−1; ΔG o=−6.0±0.3 kcal/mol).The ΔG o of the strong and weak complexes was independent of the solution ionic strength, 0.0022≤μ≤0.022. A model of EtBr binding with single-stranded regions of poly(dA)poly(dT) is discussed.  相似文献   

13.
14.
The interaction of bioactive protoberberine alkaloids berberine, palmatine, and coralyne with the DNA triplex poly(dT)⋅(poly(dA)⋅poly(dT)) was studied using biophysical and calorimetric techniques. All three alkaloids bound the triplex cooperatively. Berberine and palmatine predominantly stabilized the triplex structure, while coralyne stabilized both triplex and duplex structures as inferred from optical thermal melting profiles. Fluorescence quenching, polarization, and viscometric studies hinted at an intercalative mode of binding for the alkaloids to the triplex, coralyne being more strongly intercalated compared to partial intercalation of berberine and palmatine. The overall affinity of coralyne was two order higher (2.29×107 M −1) than that of berberine (3.43×105 M −1) and palmatine (2.34×105 M −1). Isothermal titration calorimetric studies revealed that the binding to the triplex was favored by negative enthalpy change (ΔH=−3.34 kcal/mol) with favorable entropy contribution (TΔS = 4.07 kcal/mol) for berberine, favored by almost equal negative enthalpy (ΔH =−3.88 kcal/mol) and entropy changes (TΔS = 3.37 kcal/mol) for palmatine, but driven by large enthalpy contributions (ΔH =−25.62 kcal/mol and TΔS =−15.21 kcal/mol) for coralyne. These results provide new insights on the binding of isoquinoline alkaloids to the DNA triplex structure.  相似文献   

15.
16.
We report the temperature and salt dependence of the volume change (DeltaVb) associated with the binding of ethidium bromide and netropsin with poly(dA).poly(dT) and poly[d(A-T)].poly[d(A-T)]. The DeltaV(b) of binding of ethidium with poly(dA).poly(dT) was much more negative at temperatures approximately 70 degrees C than at 25 degrees C, whereas the difference is much smaller in the case of binding with poly[d(A-T)].poly[d(A-T)]. We also determined the volume change of DNA-drug interaction by comparing the volume change of melting of DNA duplex and DNA-drug complex. The DNA-drug complexes display helix-coil transition temperatures (Tm several degrees above those of the unbound polymers, e.g., the Tm of the netropsin complex with poly(dA)poly(dT) is 106 degrees C. The results for the binding of ethidium with poly[d(A-T)].poly[d(A-T)] were accurately described by scaled particle theory. However, this analysis did not yield results consistent with our data for ethidium binding with poly(dA).poly(dT). We hypothesize that heat-induced changes in conformation and hydration of this polymer are responsible for this behavior. The volumetric properties of poly(dA).poly(dT) become similar to those of poly[d(A-T)].poly[d(A-T)] at higher temperatures.  相似文献   

17.
18.
Mercuric binding studies at pH 10 revealed that poly(dA): poly(dT) exhibits a more dramatic absorption spectral alteration than the alternating polymer poly(dA-dT):poly(dA-dT) and induces a unique intense positive CD band at 296 nm during the spectral titrations. Comparative studies with its component single strands suggest that the spectral alterations exhibited by poly(dA): poly(dT) are consistent with a binding model in which the mercuric ions initially bind to thymines and cause the eventual strand separation of the duplex, with subsequent high cooperative binding to the poly(dA) strands. This interpretation is supported by the binding isotherms indicating much stronger mercuric binding to poly(dT) than to poly(dA), with saturation binding densities of 1 Hg(II) per 2 bases and 1 Hg(II) per base, respectively, and very high binding cooperativity for poly(dA). Striking spectral alterations are exhibited by the mercuric binding to poly(dA), likely the consequence of binding to the amino group of dA in an alkaline solution. The mononucleoside dA exhibits minor spectral alterations upon similar mercuric chloride additions whereas the dinucleoside monophosphate d(AA) exhibits significant spectral changes, albeit less pronounced than those of poly(dA). Some sequence effects on the mercuric binding are observed in the dinucleotide studies. Our CD results on the mercuric binding to polynucleotides do not support the contention of (psi)-type condensed complex formation.  相似文献   

19.
J E Herrera  J B Chaires 《Biochemistry》1989,28(5):1993-2000
Circular dichroism and UV absorbance spectroscopy were used to monitor and characterize a premelting conformational transition of poly(dA)-poly(dT) from one helical form to another. The transition was found to be broad, with a midpoint of tm = 29.9 degrees C and delta HVH = +19.9 kcal mol-1. The transition renders poly(dA)-poly(dT) more susceptible to digestion by DNase I and facilitates binding of the intercalator daunomycin. Dimethyl sulfoxide was found to perturb poly(dA)-poly(dT) structure in a manner similar to temperature. These combined results suggest that disruption of bound water might be linked to the observed transition. A thermodynamic analysis of daunomycin binding to poly(dA)-poly(dT) shows that antibiotic binding is coupled to the polynucleotide conformational transition. Daunomycin binding renders poly(dA)-poly(dT) more susceptible to DNase I digestion at low binding ratios, in contrast to the normal behavior of intercalators, indicating that antibiotic binding alters the conformation of the polynucleotide. The unusual thermodynamic profiles previously observed for the binding of many antibiotics to poly(dA)-poly(dT) can be explained by our results as arising from the coupling of ligand binding to the polynucleotide conformational transition. Our data further suggest a physical basis for the temperature dependence of DNA bending.  相似文献   

20.
The paper presents results obtained in conformational analysis of homopolymeric four-stranded poly(dT).poly(dA).poly(dA).poly(dT) DNA helices in which the pairs of strands with identical bases are parallel and have a two-fold symmetry axis. All possible models of base binding to yield a symmetric complex have been considered. The dihedral angles of sugar-phosphate backbones and helix parameters, which are consistent with the minima of conformational energy for four-stranded DNAs, have been determined using the results of optimization of conformational energy calculated at atom-atom approximation. Potential energy is shown to depend on the structure of base complexes and on the mutual orientation of unlike strands. Possible biological functions of four-stranded helices are discussed.  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