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.
Hasegawa S  Rao J 《FEBS letters》2006,580(6):1592-1596
The internal guiding sequence (IGS) is normally located at the 5' end of trans-splicing ribozymes that are derived from the Tetrahymena group I intron, and is required for the recognition of substrate RNAs and for trans-splicing reactions. Here, we separated the Tetrahymena group I intron at the L2 loop to produce two fragments: the IGS-containing substrate, and the IGS-lacking ribozyme. We show here that two fragments can complex not through the IGS interaction but under the guidance of appended interacting nucleotides, and perform trans-splicing. The splicing reactions took place both in vitro and in mammalian cells, and the spliced mRNA product from the self-assembled ribozyme complex can be translated into functional proteins in vivo. The splicing efficiency was dependent on the length of appending nucleotides.  相似文献   

2.
Positions 2-6 of the substrate-binding internal guide sequence (IGS) of the L-21 Sca I form of the Tetrahymena thermophila intron were mutagenized to produce a GN5 IGS library. Ribozymes within the GN5 library capable of efficient cleavage of an 818-nt human immunodeficiency virus type 1 vif-vpr RNA, at 37 degrees C, were identified by ribozyme-catalyzed guanosine addition to the 3' cleavage product. Three ribozymes (IGS = GGGGCU, GGCUCC, and GUGGCU) within the GN5 library that actively cleaved the long substrate were characterized kinetically and compared to the wild-type ribozyme (GGAGGG) and two control ribozymes (GGAGUC and GGAGAU). The two control ribozymes have specific sites within the long substrate, but were not identified during screening of the library. Under single-turnover conditions, ribozymes GGGGCU, GGCUCC, and GUGGCU cleaved the 818-nt substrate 4- to 200-fold faster than control ribozymes. Short cognate substrates, which should be structureless and therefore accessible to ribozyme binding, were cleaved at similar rates by all ribozymes except GGGGCU, which showed a fourfold rate enhancement. The rate of cleavage of long relative to short substrate under single-turnover conditions suggests that GGCUCC and GUGGCU were identified because of accessibility to their specific cleavage sites within the long substrate (substrate-specific effects), whereas GGGGCU was identified because of an enhanced rate of substrate binding despite a less accessible site in the long substrate. Even though screening was performed with 100-fold excess substrate (relative to total ribozyme), the rate of multiple-turnover catalysis did not contribute to identification of trans-cleaving ribozymes in the GN5 library.  相似文献   

3.
4.
5.
Ⅰ型内含子核酶经过设计特定的信号引导序列(IGS),可特异性地定点剪接目的基因RNA,从而在RNA水平达到修复病变基因的目的。以四膜虫材料,克隆了其26S rRNA内含子核酶基因,体外转录证实该I型内含子核酶具有完全的自我剪接的功能。为检测该核酶的反式剪接功能,构建了缺失后半段564bp基因序列的绿色荧光蛋白(GFP)的截短突变体重组质粒XYQ5/XYQ10pEGFP-C-2,并证实其失去了发射绿色荧光的活性。利用PCR和分子克隆技术,构建了以上EGFP突变体的反式剪接修复核酶ptrans-rib-CMV2,该核酶载体以克隆的26S RNA内含子为核心,选择EGFP编码区194位TG为剪接位点,以188-193位设计IGS序列,核酶3′端携带195-890bp的EGFP基因序列,连接于pRC-CMV2真核表达载体中。体外转录突变EGFP的原核表达载体XYQ5/10-pGEM和ptrans-rib-CMV2,以混合转录产物为模板进行RT-PCR,电泳及测序证实产物中含有反式剪接修复的野生型EGFP mRNA,从而证实构建的反式剪接核酶具有体外反式剪接功能。将截短突变重组质粒XYQ5/XYQ10- pEGFP-C2与核酶质粒ptrans-rib-CMV2共转染Hela细胞,用荧光显微镜观察转染结果,发现有少量共转染的Hela细胞发出绿光;RT-PCR检测出野生型EGFP mRNA,证明构建的反式剪接核酶具有体内反式剪接的功能,但其反式剪接效率低。  相似文献   

6.
We have reconstituted a group I self-splicing reaction between two RNA molecules with different functional RNA parts: a substrate molecule containing the 5' splice site and a functional internal guide sequence (IGS), and a ribozyme molecule with core structure elements and splice sites but a mutated IGS. The 5' exon of the substrate molecule is ligated in trans to the 3' exon of the ribozyme molecule, suggesting that the deficient IGS in the ribozyme can be replaced by an externally added IGS present on the substrate molecule. This result is different from catalysis mediated by proteins where it is not possible to dissect the specificity of an enzyme from its catalytic activity.  相似文献   

7.
8.
9.
Zarrinkar PP  Sullenger BA 《Biochemistry》1999,38(11):3426-3432
Group I ribozymes can repair mutant RNAs via trans-splicing. Unfortunately, substrate specificity is quite low for the trans-splicing reaction catalyzed by the group I ribozyme from Tetrahymenathermophila. We have used a systematic approach based on biochemical knowledge of the function of the Tetrahymena ribozyme to optimize its ability to discriminate against nonspecific substrates in vitro. Ribozyme derivatives that combine a mutation which indirectly slows down the rate of the chemical cleavage step by weakening guanosine binding with additional mutations that weaken substrate binding have greatly enhanced specificity with short oligonucleotide substrates and an mRNA fragment derived from the p53 gene. Moreover, compared to the wild-type ribozyme, reaction of a more specific ribozyme with targeted substrates is much less sensitive to the presence of nonspecific RNA competitors. These results demonstrate how a detailed understanding of the biochemistry of a catalytic RNA can facilitate the design of customized ribozymes with improved properties for therapeutic applications.  相似文献   

10.
Group I intron ribozymes can repair mutated mRNAs by replacing the 3′-terminal portion of the mRNA with their own 3′-exon. This trans-splicing reaction has the potential to treat genetic disorders and to selectively kill cancer cells or virus-infected cells. However, these ribozymes have not yet been used in therapy, partially due to a low in vivo trans-splicing efficiency. Previous strategies to improve the trans-splicing efficiencies focused on designing and testing individual ribozyme constructs. Here we describe a method that selects the most efficient ribozymes from millions of ribozyme variants. This method uses an in vivo rescue assay where the mRNA of an inactivated antibiotic resistance gene is repaired by trans-splicing group I intron ribozymes. Bacterial cells that express efficient trans-splicing ribozymes are able to grow on medium containing the antibiotic chloramphenicol. We randomized a 5′-terminal sequence of the Tetrahymena thermophila group I intron and screened a library with 9 × 106 ribozyme variants for the best trans-splicing activity. The resulting ribozymes showed increased trans-splicing efficiency and help the design of efficient trans-splicing ribozymes for different sequence contexts. This in vivo selection method can now be used to optimize any sequence in trans-splicing ribozymes.  相似文献   

11.
Interactions of the 5' and 3' splice sites with intron internal sequences of the yeast mitochondrial group I intron aI3 were studied using mutation analysis. The results can be fully explained by the splice guide model in which the splice sites are defined by the Internal Guide Sequence. No evidence was found for an alternative interaction between intron nucleotides preceding the 3' splice site and nucleotides in the vicinity of the core region as was found for the Tetrahymena intron. Our results also suggest that binding of the 5' and 3' splice site nucleotides to the IGS can not take place simultaneously. The intron must therefore undergo conformational changes as the reaction proceeds from the first step of self splicing, GTP attack at the 5' splice site, to exon ligation, the second step.  相似文献   

12.
13.
14.
mRNA maturation in Trypanosoma brucei depends upon trans splicing, and variations in trans-splicing efficiency could be an important step in controlling the levels of individual mRNAs. RNA splicing requires specific sequence elements, including conserved 5' splice sites, branch points, pyrimidine-rich regions [poly(Y) tracts], 3' splice sites (3'SS), and sometimes enhancer elements. To analyze sequence requirements for efficient trans splicing in the poly(Y) tract and around the 3'SS, we constructed a luciferase-beta-galactosidase double-reporter system. By testing approximately 90 sequences, we demonstrated that the optimum poly(Y) tract length is approximately 25 nucleotides. Interspersing a purely uridine-containing poly(Y) tract with cytidine resulted in increased trans-splicing efficiency, whereas purines led to a large decrease. The position of the poly(Y) tract relative to the 3'SS is important, and an AC dinucleotide at positions -3 and -4 can lead to a 20-fold decrease in trans splicing. However, efficient trans splicing can be restored by inserting a second AG dinucleotide downstream, which does not function as a splice site but may aid in recruitment of the splicing machinery. These findings should assist in the development of improved algorithms for computationally identifying a 3'SS and help to discriminate noncoding open reading frames from true genes in current efforts to annotate the T. brucei genome.  相似文献   

15.
Dème E  Nolte A  Jacquier A 《Biochemistry》1999,38(10):3157-3167
The splicing process catalyzed by group II intron ribozymes follows the same two-step pathway as nuclear pre-mRNA splicing. In vivo, the first splicing step of wild-type introns is a transesterification reaction giving rise to a branched lariat intron-3'-exon intermediate characteristic of this splicing mode. In the wild-type introns, the ribozyme core and the substrate intron-exon junctions are carried by the same precursor molecule, making it difficult to distinguish between RNA folding and catalysis under normal splicing reactions. To characterize the catalytic step of the first transesterification reaction, we studied the reversal of this reaction, reverse branching. In this reverse reaction, the excised lariat intron and the substrate 5'-exon can be preincubated and folded separately, allowing the measure of the catalytic rate of the reaction. To measure the catalytic rate of the second splicing step, purified lariat intron-3'-exon intermediate molecules were preincubated and folded prior to the addition of 5'-exon. Conditions could be found where chemistry appeared rate limiting for both catalytic steps. Study of the metal ion requirements under these conditions resulted in the unexpected finding that, for the intron studied, substitution of magnesium ions by manganese ions enhanced the rate of the first transesterification reaction by two orders of magnitude but had virtually no effect on the second transesterification reaction or the 5' splice site cleavage by hydrolysis. Finally, the catalytic rates measured under optimal conditions for both splicing steps were faster by three orders of magnitude in the branching pathway than in the hydrolytic pathway.  相似文献   

16.
In Caenorhabditis elegans, pre-mRNAs of many genes are trans-spliced to one of two spliced leaders, SL1 or SL2. Some of those that receive exclusively SL1 have been characterized as having at their 5' ends outrons, AU-rich sequences similar to introns followed by conventional 3' splice sites. Comparison of outrons from many different SL1-specific C. elegans genes has not revealed the presence of any consensus sequence that might encode SL1-specificity. In order to determine what parameters influence the splicing of SL1, we performed in vivo experiments with synthetic splice sites. Synthetic AU-rich RNA, 51 nt or longer, placed upstream of a consensus 3' splice site resulted in efficient trans-splicing. With all sequences tested, this trans-splicing was specifically to SL1. Thus, no information beyond the presence of AU-rich RNA at least as long as the minimum-length C. elegans intron, followed by a 3' splice site, is required to specify trans-splicing or for strict SL1 specificity.  相似文献   

17.
Spliceosome-mediated RNA trans-splicing (SMaRT) has been used previously to reprogram mutant endogenous CFTR and factor VIII mRNAs in human epithelial cell and tissue models and knockout mice, respectively. Those studies used 3' exon replacement (3'ER); a process in which the distal portion of RNA is reprogrammed. Here, we also show that the 5' end of mRNA can be completely rewritten by 5'ER. For proof-of-concept, and to test whether 5'ER could generate functional CFTR, we generated a mutant minigene target containing CFTR exons 10-24 (deltaF508) and a mini-intron 10, and a pretrans-splicing molecule (targeted to intron 10) containing CFTR exons 1-10 (+F508), and tested these two constructs in 293T cells for anion efflux transport. Cells cotransfected with target and PTM showed a consistent increase in anion efflux, but there was no response in control cells that received PTM or target alone. Using a LacZ reporter system to accurately quantify trans-splicing efficiency, we tested several unique PTM designs. These studies provided two important findings as follows: (1) efficient trans-splicing can be achieved by binding the PTM to different locations in the target, and (2) relatively few changes in PTM design can have a profound impact on trans-splicing activity. Tethering the PTM close to the target 3' splice site (as opposed to the donor site) and inserting an intron in the PTM coding resulted in a 65-fold enhancement of LacZ activity. These studies demonstrate that (1) SMaRT can be used to reprogram the 5' end of mRNA, and (2) efficiency can be improved substantially.  相似文献   

18.
RNA reprogramming represents a new concept in correcting genetic defects at the RNA level. However, for the technique to be useful for therapy, the level of reprogramming must be appropriate. To improve the efficiency of group I ribozyme-mediated RNA reprogramming, when using the Tetrahymena ribozyme, regions complementary to the target RNA have previously been extended in length and accessible sites in the target RNAs have been identified. As an alternative to the Tetrahymena model ribozyme, the DiGIR2 group I ribozyme, derived from a mobile group I intron in rDNA of the myxomycete Didymium iridis, represents a new and attractive tool in RNA reprogramming. We reported recently that the deletion of a structural element within the P9 domain of DiGIR2 turns off hydrolysis at the 3' splice site (side reaction) without affecting self-splicing [Haugen, P., Andreassen, M., Birgisdottir, A.B. & Johansen, S.D. (2004) Eur. J. Biochem. 271, 1015-1024]. Here we analyze the potential of the modified ribozyme, deficient in hydrolysis at the 3' splice site, for application in group I ribozyme-mediated trans-splicing of RNA. The improved ribozyme catalyses both cis-splicing and trans-splicing in vitro of a human glycosylasparaginase mRNA sequence with the same efficiency as the original DiGIR2 ribozyme, but without detectable levels of the unwanted hydrolysis.  相似文献   

19.
Experimental evidence for RNA trans-splicing in mammalian cells.   总被引:8,自引:0,他引:8       下载免费PDF全文
J Eul  M Graessmann    A Graessmann 《The EMBO journal》1995,14(13):3226-3235
  相似文献   

20.
trans-Splicing is essential for mRNA maturation in trypanosomatids. A conserved AG dinucleotide serves as the 3' splice acceptor site, and analysis of native processing sites suggests that selection of this site is determined according to a 5'-3' scanning model. A series of stable gene replacement lines were generated that carried point mutations at or near the 3' splice site within the intergenic region separating CUB2.65, the calmodulin-ubiquitin associated gene, and FUS1, the ubiquitin fusion gene of Trypanosoma cruzi. In one stable line, the elimination of the native 3' splice acceptor site led to the accumulation of Y-branched splicing intermediates, which served as templates for mapping the first trans-splicing branch points in T. cruzi. In other lines, point mutations shifted the position of the first consensus AG dinucleotide either upstream or downstream of the wild-type 3' splice acceptor site in this intergenic region. Consistent with the scanning model, the first AG dinucleotide downstream of the branch points was used as the predominant 3' splice acceptor site. In all of the stable lines, the point mutations affected splicing efficiency in this region.  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