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 46 毫秒
1.
J L Betz  M Z Fall 《Gene》1988,67(2):147-158
The specific binding of dominant-negative (I-d) lactose (lac) repressors to wild-type (wt) as well as mutant (Oc) lac operators has been examined to explore the sequence-specific interaction of the lac repressor with its target. Mutant lacI genes encoding substitutions in the N-terminal 60 amino acids (aa) were cloned in a derivative of plasmid pBR322. Twelve of these lacI-d missense mutations were transferred from F'lac episomes using general genetic recombination and molecular cloning, and nine lacI missense mutations were recloned from M13-lacI phages [Mott et al., Nucl. Acids Res. 12 (1984) 4139-4152]. The mutant repressors were examined for polypeptide size and stability, for binding the inducer isopropyl-beta-D-thiogalactoside (IPTG), as well as binding to wt operator. The mutant repressors' affinities for wt operator ranged from undetectable to about 1% that of wt repressor, and the mutant repressors varied in transdominance against repressor expressed from a chromosomal lacIq gene. Six of the I-d repressors were partially degraded in vivo. All repressors bound IPTG with approximately the affinity of wt repressor. Repressors having significant affinity for wt operator or with substitutions in the presumed operator recognition helix (aa 17-25) were examined in vivo for their affinities for a series of single site Oc operators. Whereas the Gly-18-, Ser-18- and Leu-18-substituted repressors showed altered specificity for position 7 of the operator [Ebright, Proc. Natl. Acad. Sci. USA 83 (1986) 303-307], the His-18 repressor did not affect specificity. This result may be related to the greater side-chain length of histidine compared to the other amino acid substitutions.  相似文献   

2.
Sequence similarities among monkey ori-enriched (ors) fragments   总被引:4,自引:0,他引:4  
Nucleotide sequences have been determined for eight ors (ori-enriched sequence) fragments isolated from monkey DNA by a method that was designed to enrich for origins of DNA replication [Kaufmann et al., Mol. Cell. Biol. 5 (1985) 721-727]. Evidence has been presented that some or possibly all of these sequences can serve, albeit inefficiently, as oris in vivo [Frappier and Zannis-Hadjopoulos, Proc. Natl. Acad. Sci. USA 84 (1987) 6668-6672]. Two of the fragments were found to contain the long terminal repeat-like elements of the 'O-family' of moderately repetitive sequences that are present in human DNA as a transposon-like element [Paulson et al., Nature 315 (1985) 359-361]. Extensive pair-wise comparisons of the sequences failed to detect any statistically significant common sequences, except for long asymmetrically distributed A + T-rich stretches. Nonetheless, when the ors fragments were examined for the presence of published consensus sequences, seven of eight were found to contain the control sequence described by Dierks et al. [Cell 32 (1983) 695-706], and the same seven of eight were found to contain both the scaffold attachment region T consensus [Gasser and Laemmli, Cell 46 (1986) 521-530] and the minimal Saccharomyces cerevisiae autonomously replicating sequence consensus [e.g., Palzkill and Newlon, Cell 53 (1988) 441-450].  相似文献   

3.
4.
The nature and dynamics of the motions of a diunsaturated fatty acyl chain in a lipid bilayer were examined using a comprehensive simulation program for 2H NMR line shapes developed by Wittebort et al. [Wittebort, R. J., Olejniczak, E. T., & Griffin, R. G. (1987) J. Chem. Phys. 36, 5411-5420]. A motional model in which the isolinoleoyl chain (18:2 delta 6,9) adopts two conformations consistent with the low energy structures proposed for 1,4-pentadiene [Applegate, K. R., & Glomset, J. A. (1986) J. Lipid Res. 27, 658-680], but undergoes a rapid jump between these states, is sufficient to account for the experimentally observed quadrupolar couplings, the 2H-2H and 1H-2H dipolar couplings, the longitudinal relaxation times, and the changes in the average conformation of the chain that occur with a variation in temperature. The jump motion originates via rotations about the C7-C8 and the C8-C9 carbon bonds and leads to the low order parameters assigned to the C8 methylene segment (0.18) and the C9-C10 double bond (0.11). In contrast, the C6-C7 double bond, which is not involved in the two-site jump, characterized by a relatively large order parameter (0.56). Fatty acyl chains containing three or more double bonds likely cannot undergo the same jump motion and consequently will be highly ordered structures. Correlation times for diffusion of the molecular long axis of the diunsaturated acyl chain about the bilayer normal (approximately 10(-10) s) and for the local jump motion (approximately 10(-10) s) were calculated.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
The 33-kDa manganese-stabilizing protein (MSP) of Photosystem II (PS II) maintains the functional stability of the Mn cluster in the enzyme’s active site. This protein has been shown to possess characteristics similar to those of the intrinsically disordered, or natively unfolded proteins [Lydakis-Simantiris et al. (1999b) Biochemistry 38: 404–414]. Alternately it was proposed that MSP should be classified as a molten globule, based in part on the hypothesis that its lone disulfide bridge is necessary for structural stability and function in solution [Shutova et al. (2000) FEBS Lett. 467: 137–140]. A site-directed mutant MSP (C28A,C51A) that eliminates the disulfide bond reconstitutes O2 evolution activity and binds to MSP-free PS II preparations at wild-type levels [Betts et al. (1996) Biochim. Biophys. Acta 1274: 135–142]. This mutant was further characterized by incubation at 90 °C to determine the effect of loss of the disulfide bridge on MSP thermostability and solution structure. After heating at 90 °C for 20 min, C28A,C51A MSP was still able to bind to PS II preparations at molar stoichiometries similar to those of WT MSP and reconstitute O2 evolution activity. A fraction of the protein aggregates upon heating, but after resolubilization, it regains the ability to bind to PS II and reconstitute O2 evolution activity. Characterization of the solution structure of C28A,C51A MSP, using CD spectroscopy, UV absorption spectroscopy, and gel filtration chromatography, revealed that the mutant has a more disordered solution structure than WT MSP. The disulfide bond is therefore unnecessary for MSP function and the intrinsically disordered characteristics of MSP are not dependent on its presence. However, the disulfide bond does play a role in the solution structure of MSP in vivo, as evidenced by the lability of a C20S MSP mutation in Synechocystis 6803 [Burnap et al. (1994) Biochemistry 33: 13712–13718].  相似文献   

6.
The introduction of a disulfide bond into the neutral protease from Bacillus stearothermophilus by the double mutation G8C/N60C had resulted in an extremely thermostable enzyme with a half-life of 35.9 min at 92.5 degrees C [Mansfeld, J., Vriend, G., Dijkstra, B.W., Veltman, O.R., van den Burg, B., Venema, G., Ulbrich-Hofmann, R. & Eijsink, V.G. (1997) J. Biol. Chem. 272, 11152-11156]. The study in guanidine hydrochloride of this enzyme and the respective wild-type enzyme allowed us to distinguish between the stability toward global unfolding and autoproteolysis. At low protease concentrations (20 microg.mL-1) and short periods of incubation with guanidine hydrochloride (5 min), transition curves without the interference by autoproteolysis could be derived from fluorescence emission measurements. The effect of the disulfide bond on the global unfolding of the protein proved to be smaller than expected. In contrast, the measurement of autoproteolysis at higher protein concentrations (100 microg.mL-1) by quantitative evaluation of the bands of intact protein on SDS/PAGE revealed a strong stabilization toward autoproteolytic degradation by the disulfide bond. The rate of autoproteolysis in guanidine hydrochloride was found to be much lower than that of thermal denaturation, which can be attributed to the inhibition of the proteases by this denaturant. The results suggest that the disulfide bond stabilizes the protease against autoproteolysis more than against global unfolding. Autoproteolysis starts as soon as the cleavage sites in flexible external structural regions become accessible. It is suggested that the stabilizing effect of the disulfide bond is caused by the fixation of the crucial loop region 56-69 or by hindrance of the primary cleavage in this region by the amino acid exchanges.  相似文献   

7.
Antithrombin (AT) is a major plasma protease inhibitor with three intramolecular disulfide bonds and a deficiency of it is associated with venous thrombosis. Recently, we prepared CHO cells overexpressing a novel mutant, AT(C95R), with a disulfide bond removed, and revealed that this mutant remained for a long time in the endoplasmic reticulum (ER) without being degraded and also accumulated in newly formed membrane structures that resembled Russell bodies (RB) [Tanaka, Y. et al. (2002) J. Biol. Chem. 277, 51058-51067]. In this study, we replaced each of the individual cysteine residues of AT with an arginine and also two paired cysteine residues with arginines. We stably expressed these mutant ATs in CHO cells, and examined the roles of each cysteine residue or disulfide bond in the accumulation of mutant ATs and the formation of RB-like structures. In pulse-chase experiments, the secretion of mutant ATs with single mutations decreased approximately 1/5-1/50 times compared to that of the wild type AT. All of the mutant ATs were retained in the ER and were also found to accumulate in the RB-like structures. On the other hand, the fates of mutant ATs with double mutations fell into two categories. Secretion of mutant AT(C8R,C128R) decreased only approximately 1/2 times and no RB-like structures appeared. Mutants AT(C21R,C95R) and AT(C247R,C430R) exhibited similar secretion kinetics to the mutant ATs with the single mutations and were found in RB-like structures. On a sucrose gradient, all of the mutant ATs that induced RB-like structures migrated as oligomeric structures, whereas wild type AT and AT(C8R,C128R) migrated as monomers. Further, to clarify the morphological pathway through which RB-like structures are formed, we prepared CHO cells in which the expression of AT(C95R) was controlled by the Tet-On system. During expression of AT(C95R), RB-like structures formed through expansion of the ER. These results suggest that the correct folding with each disulfide bond is essential for the secretion of AT and oligomerization of mutant ATs in the ER is involved in the formation of RB-like structures.  相似文献   

8.
According to the X-ray structure analysis of an EcoRI-oligodeoxynucleotide complex [McClarin et al. (1986) Science 234, 1526], sequence specificity is mediated by 12 hydrogen bonds, 6 from each of the two identical subunits of the dimeric enzyme to the recognition site -GAATTC-: Arg200 forms two hydrogen bonds with guanine, while Glu144 and Arg145 form four hydrogen bonds to adjacent adenine residues. Changing the hydrogen-bonding potential at the recognition site without perturbing the rest of the interface should lead to the recognition of degenerate sequences [Rosenberg et al. (1987) in Protein Engineering (Oxender, D. L., & Fox, C. F., Eds.) pp 237-250, Liss, New York]. We have shown previously that replacing Glu144 by Gln and Arg145 by Lys affects the activity of the enzyme, not, however, its specificity [Wolfes et al. (1986) Nucleic Acids Res. 14, 9063]. We show now that also the mutation of Arg200 to Lys, the double mutation Glu144Arg145 to GlnLys, and the triple mutation Glu144Arg145Arg200 to GlnLysLys do not lead to a detectable degeneracy of the specificity of cleavage by EcoRI but significantly impair the catalytic activity of this enzyme. A detailed analysis of the steady-state kinetics of cleavage of pUC8 DNA and a tridecadeoxynucleotide substrate demonstrates that the reduction in activity for all DNA binding site mutants investigated so far is mainly due to a decrease in kcat, with the exception of the Arg200 to Lys mutant, which is only impaired in its KM.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
A 24,000 Dalton protein with follicle stimulating hormone (FSH)-releasing activity, named activin, has been characterized previously from porcine follicular fluid as a heterodimer composed of the beta-subunits of inhibins A and B linked by disulfide bond(s) [Ling et al. (1986) Nature, in press]. In this paper we report the isolation of another 24,000 Dalton protein with FSH-releasing activity from porcine follicular fluid, using successive steps of heparin-Sepharose affinity chromatography, gel filtration on Sephacryl S-200, and four steps of reversed-phase HPLC, followed by preparative sodium dodecyl-sulfate-polyacrylamide gel electrophoresis chromatography. Based on the molecular weight of the isolated molecule and its deduced NH2-terminal sequence, we propose that this second FSH-releasing substance present in porcine follicular fluid is a homodimeric protein composed of two beta-subunits of inhibin A joined together by disulfide bond(s). The name homo-activin-A is proposed for this substance.  相似文献   

10.
Daunorubicin, an anthracycline antitumor antibiotic, can be complexed in egg white apo-riboflavin binding protein. The reduction of this complex was studied by gamma-radiolysis and pulse radiolysis using COO.- free radicals as reductants. The final products are 7-deoxydaunomycinone intercalated in the protein and thiol groups coming from the reduction of disulfide bonds of the protein, in the respective proportions of 90% and 10%. One-electron reduction of the complex gives daunorubicin semiquinone radical and a disulfide anion. The rate constants of the reactions of COO.- ions with the complex and with the disulfide bond in the protein alone are respectively equal to 2.4 x 10(8) mol-1.L.s-1 and 6.4 x 10(7) mol-1.L.s-1. Daunorubicin semiquinone decays by a first-order process, the rate constant of which is independent of the initial protein and radical concentrations. Without protein, daunorubicin semiquinone undergoes a disproportionation-comproportionation equilibrium [Houée-Levin, C., Gardès-Albert, M., Ferradini, C., Faraggi, M., & Klapper, M. (1985) FEBS Lett. 179, 46-50]. We propose that a protein residue reduces semiquinone by an intramolecular path. This creates an electron hole in the protein which may alter its function. This reduction process is very different from the reduction mechanism of riboflavin binding protein by the same reductant [Faraggi, M., Steiner, J.P., & Klapper, M.H. (1985) Biochemistry 24, 3273-3279]. These results suggest a new deleterious pathway to explain the antitumor and/or cytotoxic effect of this drug.  相似文献   

11.
H Kikuchi  Y Goto  K Hamaguchi 《Biochemistry》1986,25(8):2009-2013
The constant (CL) fragment of the immunoglobulin light chain contains only one intrachain disulfide bond buried in the interior of the molecule. The kinetics of reduction with dithiothreitol of the disulfide bond were studied at various concentrations of guanidine hydrochloride at pH 8.0 and 25 degrees C. It was found that the disulfide bond is reduced even in the absence of guanidine hydrochloride. The results of the reduction kinetics were compared with those of the unfolding and refolding kinetics of the CL fragment previously reported [Goto, Y., & Hamaguchi, K. (1982) J. Mol. Biol. 156, 891-910]. It was shown that the reduction of the disulfide bond proceeds through a species with a conformation very similar to that of the fully unfolded one and that the CL fragment undergoes global unfolding transition even in water.  相似文献   

12.
Lee YH  Tanner JJ  Larson JD  Henzl MT 《Biochemistry》2004,43(31):10008-10017
In model peptide systems, Ca2+ affinity is maximized in EF-hand motifs containing four carboxylates positioned on the +x and -x and +z and -z axes; introduction of a fifth carboxylate ligand reduces the affinity. However, in rat beta-parvalbumin, replacement of Ser-55 with aspartate heightens divalent ion affinity [Henzl, M. T., et al. (1996) Biochemistry 35, 5856-5869]. The corresponding alpha-parvalbumin variant (S55D/E59D) likewise exhibits elevated affinity [Henzl, M. T., et al. (2003) Anal. Biochem. 319, 216-233]. To determine whether these mutations produce a variation on the archetypal EF-hand coordination scheme, we have obtained high-resolution X-ray crystallographic data for alpha S55D/E59D. As anticipated, the aspartyl carboxylate replaces the serine hydroxyl at the +z coordination position. Interestingly, the Asp-59 carboxylate abandons the role it plays as an outer sphere ligand in wild-type rat beta, rotating away from the Ca2+ and, instead, forming a hydrogen bond with the amide of Glu-62. Superficially, the coordination sphere in the CD site of alpha S55D/E59D resembles that in the EF site. However, the orientation of the Asp-59 side chain is predicted to stabilize the D-helix, which may contribute to the heightened divalent ion affinity. DSC data indicate that the alpha S55D/E59D variant retains the capacity to bind 1 equiv of Na+. Consistent with this finding, when binding measurements are conducted in K(+)-containing buffer, divalent ion affinity is markedly higher. In 0.15 M KCl and 0.025 M Hepes-KOH (pH 7.4) at 5 degrees C, the macroscopic Ca2+ binding constants are 1.8 x 10(10) and 2.0 x 10(9) M(-1). The corresponding Mg2+ binding constants are 2.7 x 10(6) and 1.2 x 10(5) M(-1).  相似文献   

13.
The multi-dynein hypothesis [Asai, 1995: Cell Motil Cytoskeleton 32:129-132] states: (1) there are many different dynein HC isoforms; (2) each isoform is encoded by a different gene; (3) different isoforms have different functions. Many studies provide evidence in support of the first two statements [Piperno et al., 1990: J Cell Biol 110:379-389; Kagami and Kamiya, 1992: J Cell Sci 103:653-664; Gibbons, 1995: Cell Motil Cytoskeleton 32:136-144; Porter et al., 1996: Genetics 144:569-585; Xu et al., 1999: J Eukaryot Microbiol 46:606-611] and there is evidence that outer arms and inner arms play different roles in flagellar beating [Brokaw and Kamiya, 1987: Cell Motil. Cytoskeleton 8:68-75]. However, there are few studies rigorously testing in vivo whether inner arm dyneins, especially the 1-headed inner arm dyneins, play unique roles. This study tested the third tenet of the multi-dynein hypothesis by introducing mutations into three inner arm dynein HC genes (DYH8, 9 and 12) that are thought to encode HCs associated with 1-headed inner arm dyneins. Southern blots, Northern blots, and RT-PCR analyses indicate that all three mutants (KO-8, 9, and 12) are complete knockouts. Each mutant swims slower than the wild-type cells. The beat frequency of KO-8 cells is lower than that of the wild-type cells while the beat frequencies of KO-9 and KO-12 are not different from that of wild-type cells. Our results suggest that each inner arm dynein HC is essential for normal cell motility and cannot be replaced functionally by other dynein HCs and that not all of the 1-headed inner arm dyneins play the same role in ciliary motility. Thus, the results of our study support the multi-dynein hypothesis [Asai, 1995: Cell Motil Cytoskeleton 32:129-132].  相似文献   

14.
We have used an alkaline phosphatase protection assay to investigate the interaction of the trp repressor with its operator sequence. The assay is based on the principle that the trp repressor will protect a terminally 5'-32P-labeled operator DNA fragment from attack by alkaline phosphatase. The optimal oligonucleotide for investigating the trp repressor/operator interaction extends two base pairs from each end of the genetically defined target sequence predicted by in vivo studies [Bass et al. (1987) Genes Dev. 1, 565-572]. The assay works well over a 10,000-fold range of protein/DNA affinity and is used to show that the corepressor, L-tryptophan, causes the liganded repressor to bind a 20 base pair trp operator duplex 6400 times more strongly than the unliganded aporepressor. The affinity of the trp repressor for operators containing symmetrical mutations was interpreted in terms of the trp repressor/operator crystal structure as follows: (1) Direct hydrogen bonds with the functional groups of G-9 of the trp operator and the side chain of Arg 69 of the trp repressor contribute to DNA-binding specificity. (2) G-6 of the trp operator is critical for DNA-binding specificity probably because of the two water-mediated hydrogen bonds between its functional groups and the N-terminus of the trp repressor's E-helix. (3) Sequence-dependent aspects of the trp operator's conformation help stabilize the trp repressor/operator complex.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
Li SX  Vaccaro JA  Sweasy JB 《Biochemistry》1999,38(15):4800-4808
DNA polymerase beta is a small monomeric polymerase that participates in base excision repair and meiosis [Sobol, R., et al. (1996) Nature 379, 183-186; Plug, A., et al. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 1327-1331]. A DNA polymerase beta mutator mutant, F272L, was identified by an in vivo genetic screen [Washington, S., et al. (1997) Proc. Natl. Acad. Sci. U.S.A. 94, 1321-1326]. Residue 272 is located within the deoxynucleoside triphosphate (dNTP) binding pocket of DNA polymerase beta according to the known DNA polymerase beta crystal structures [Pelletier, H., et al. (1994) Science 264, 1891-1893; Sawaya, M., et al. (1997) Biochemistry 36, 11205-11215]. The F272L mutant produces errors at a frequency 10-fold higher than that of wild type in vivo and in the in vitro HSV-tk gap-filling assay. F272L shows an increase in the frequency of both base substitution mutations and frameshift mutations. Single-enzyme turnover studies of misincorporation by wild type and F272L DNA polymerase beta demonstrate that there is a 4-fold decrease in fidelity of the mutant as compared to that of the wild type enzyme for a G:A mismatch. The decreased fidelity is due primarily to decreased discrimination between the correct and incorrect dNTP during ground-state binding. These results suggest that the phenylalanine 272 residue is critical for maintaining fidelity during the binding of the dNTP.  相似文献   

16.
17.
Laity JH  Montelione GT  Scheraga HA 《Biochemistry》1999,38(50):16432-16442
We have identified specific regions of the polypeptide chain of bovine pancreatic ribonuclease A (RNase A) that are critical for stabilizing the oxidative folding intermediate des-[40-95] (with three native disulfide bonds but lacking the fourth native Cys40-Cys95 disulfide bond) in an ensemble of largely disordered three-disulfide precursors (3S if des-[40-95]). A stable analogue of des-[40-95], viz., [C40A, C95A] RNase A, which contains three out of four native disulfide pairings, was previously found to have a three-dimensional structure very similar to that of the wild-type protein. However, it is determined here from GdnHCl denaturation experiments to have significantly reduced global stability, i.e., = 4.5 kcal /mol at 20 degrees C and pH 4.6. The local stability of [C40A, C95A] RNase A was also examined using site-specific amide (2)H/(1)H exchange measurements at pD 5.0 to determine the individual unfolding free energy of specific residues under both strongly native (12 degrees C) and more destabilizing (20 degrees C) conditions. Comparison of the relative stabilities at specific amide sites of [C40A, C95A] RNase A at both temperatures with the corresponding values for the wild-type protein at 35 degrees C corroborates previous experimental evidence that unidentified intramolecular contacts in the vicinity of the preferentially formed native one-disulfide (C65-C72) loop are crucial for stabilizing early folding intermediates, leading to des-[40-95]. Moreover, values of for residues at or near the third alpha-helix, and in part of the second beta-sheet of [C40A, C95A] RNase A, indicate that these two regions of regular backbone structure contribute to stabilizing the global chain fold of the des-[40-95] disulfide-folding intermediate in the wild-type protein. More significantly, we have identified numerous specific residues in the first alpha-helix and the first beta-sheet of the protein that are stabilized in the final step of the major oxidative regeneration pathway of RNase A (des-[40-95] --> N).  相似文献   

18.
Induction of conductance heterogeneity in gramicidin channels   总被引:8,自引:0,他引:8  
In previous work from our laboratory, 5-10% of the channels formed by [Val1]gramicidin A have conductances that fall outside the narrow range that conventionally has defined the standard gramicidin channel [e.g., see Russell et al. (1986) Biophys. J. 49, 673]. Reports from other laboratories, however, show that up to 50% of [Val1]gramicidin channels have conductances that fall outside the range for standard channels [e.g., see Prasad et al. (1986) Biochemistry 25, 456]. This laboratory-to-laboratory variation in the distribution of gramicidin single-channel conductances suggests that the conductance variants are induced by some environmental factor(s) [Busath et al. (1987) Biophys. J. 51, 79]. In order to test whether extrinsic agents can induce such conductance heterogeneity, we examined the effects of nonionic or zwitterionic detergents upon gramicidin channel behavior. In phospholipid bilayers, detergent addition induces many changes in gramicidin channel behavior: all detergents tested increase the channel appearance rate and average duration; most detergents decrease the conductance of the standard channel; and all but one of the detergents increase the conductance heterogeneity. These results show that the conductance heterogeneity can result from environmental perturbations, thus providing a possible explanation for the laboratory-to-laboratory variation in the heterogeneity of gramicidin channels. In addition, the differential detergent effects suggest possible mechanisms by which detergents can induce the conformational perturbations that result in gramicidin single-channel conductance variations.  相似文献   

19.
Quantification of random mutations in the mitochondrial genome   总被引:1,自引:0,他引:1  
Mitochondrial DNA (mtDNA) mutations contribute to the pathology of a number of age-related disorders, including Parkinson disease [A. Bender et al., Nat. Genet. 38 (2006) 515,Y. Kraytsberg et al., Nat. Genet. 38 (2006) 518], muscle-wasting [J. Wanagat, Z. Cao, P. Pathare, J.M. Aiken, FASEB J. 15 (2001) 322], and the metastatic potential of cancers [K. Ishikawa et al., Science 320 (2008) 661]. The impact of mitochondrial DNA mutations on a wide variety of human diseases has made it increasingly important to understand the mechanisms that drive mitochondrial mutagenesis. In order to provide new insight into the etiology and natural history of mtDNA mutations, we have developed an assay that can detect mitochondrial mutations in a variety of tissues and experimental settings [M. Vermulst et al., Nat. Genet. 40 (2008) 4, M. Vermulst et al., Nat. Genet. 39 (2007) 540]. This methodology, termed the Random Mutation Capture assay, relies on single-molecule amplification to detect rare mutations among millions of wild-type bases [J.H. Bielas, L.A. Loeb, Nat. Methods 2 (2005) 285], and can be used to analyze mitochondrial mutagenesis to a single base pair level in mammals.  相似文献   

20.
We have previously described a collection of mutations in conserved residues of the core of the Tetrahymena self-splicing intron. Most of these single base substitutions have less than 10% of the activity of their parental intron derivative [Couture, S., et al., (1990) J. Mol. Biol., 215, 345-358]. We examined the effect of two agents known to stabilize RNA structure, spermidine and ethanol, on the activity of many of these mutant RNAs. In the presence of either 5 mM spermidine or 20% ethanol most substitution mutations were partially or completely suppressed. These conditions also increased the temperature optima of both wild-type and mutant ribozymes. In addition, we find that mutations are also suppressed by a high concentration of GTP, a substrate in the reaction which is bound specifically by the intron. Thus we observe a general suppression of mutations in an RNA enzyme (ribozyme) by spermidine, ethanol and by substrate stabilization. These results are consistent with the idea that most mutations destabilize the folded structure of the ribozyme and can be suppressed by any of a variety of stabilizing influences.  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