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 15 毫秒
1.
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   

2.
3.
4.
5.
Spliced-leader (SL) trans-splicing has been found in all molecularly characterized nematode species to date, and it is likely to be a nematode synapomorphy. Most information regarding SL trans-splicing has come from the study of nematodes from a single monophyletic group, the Rhabditida, all of which employ SL RNAs that are identical to, or variants of, the SL1 RNA first characterized in Caenorhabditis elegans. In contrast, the more distantly related Trichinella spiralis, belonging to the subclass Dorylaimia, utilizes a distinct set of SL RNAs that display considerable sequence diversity. To investigate whether this is true of other members of the Dorylaimia, we have characterized SL RNAs from Prionchulus punctatus. Surprisingly, this revealed the presence of a set of SLs that show clear sequence similarity to the SL2 family of spliced leaders, which have previously only been found within the rhabditine group (which includes C. elegans). Expression of one of the P. punctatus SL RNAs in C. elegans reveals that it can compete specifically with the endogenous C. elegans SL2 spliced leaders, being spliced to the pre-mRNAs derived from downstream genes in operons, but does not compete with the SL1 spliced leaders. This discovery raises the possibility that SL2-like spliced leaders were present in the last common ancestor of the nematode phylum.  相似文献   

6.
The organization of genes into operons, clusters of genes that are co-transcribed to produce polycistronic pre-mRNAs, is a trait found in a wide range of eukaryotic groups, including multiple animal phyla. Operons are present in the class Chromadorea, one of the two main nematode classes, but their distribution in the other class, the Enoplea, is not known. We have surveyed the genomes of Trichinella spiralis, Trichuris muris, and Romanomermis culicivorax and identified the first putative operons in members of the Enoplea. Consistent with the mechanism of polycistronic RNA resolution in other nematodes, the mRNAs produced by genes downstream of the first gene in the T. spiralis and T. muris operons are trans-spliced to spliced leader RNAs, and we are able to detect polycistronic RNAs derived from these operons. Importantly, a putative intercistronic region from one of these potential enoplean operons confers polycistronic processing activity when expressed as part of a chimeric operon in Caenorhabditis elegans. We find that T. spiralis genes located in operons have an increased likelihood of having operonic C. elegans homologs. However, operon structure in terms of synteny and gene content is not tightly conserved between the two taxa, consistent with models of operon evolution. We have nevertheless identified putative operons conserved between Enoplea and Chromadorea. Our data suggest that operons and “spliced leader” (SL) trans-splicing predate the radiation of the nematode phylum, an inference which is supported by the phylogenetic profile of proteins known to be involved in nematode SL trans-splicing.  相似文献   

7.
8.
9.
To cultivate the use of trans-splicing as a novel means to rapidly express various antibody fusion proteins, we tried to express antibody-reporter enzyme fusions in a COS-1 co-transfection model. When a vector designed to induce trans-splicing with IgH pre-mRNA was co-transfected with a vector encoding the mouse IgM locus, the expression of VH-secreted human placental alkaline phosphatase (SEAP) as well as Fab-SEAP were successfully expressed both in mRNA and protein levels. Especially, the vectors encoding complementary sequence to Sμ as a binding domain was accurate and efficient, producing trans-spliced mRNA of up to 2% of cis-spliced one. Since Sμ sequence should exist in every IgH pre-mRNA, our finding will lead to the rapid production and analysis of various antibody-enzyme fusions suitable for enzyme-linked immunosorbent assay (ELISA) or antibody-dependent enzyme prodrug therapy (ADEPT).  相似文献   

10.
11.
12.
Although spliceosomal introns are an abundant landmark in eukaryotic genomes, the nuclear genome of the divergent eukaryote Giardia intestinalis, the causative agent of giardiasis, has been considered as “intron-poor” with only five canonical (cis-spliced) introns. However, three research groups (including ours) have independently reported a novel class of spliceosomal introns in the G. intestinalis genome. Three protein-coding genes are split into pieces in the G. intestinalis genome, and each of the partial coding regions was independently transcribed into polyadenylated premature mRNAs (pre-mRNAs). The two pre-mRNAs directly interact with each other by an intermolecular-stem structure formed between their non-coding portions, and are then processed into mature mRNAs by spliceosome-mediated trans-splicing. Here, we summarize the recently published works on split introns (“splintrons”) in the G. intestinalis genome, and then provide our speculation on the functional property of the Giardia spliceosomes based on the putative ratio of splintrons to canonical introns. Finally, we discuss a scenario for the transition from typical GT-AG boundaries to non-typical AT-AC boundaries in a particular splintron of Giardia.  相似文献   

13.
The group I intron ribozyme from Tetrahymena was recently reengineered into a trans-splicing variant that is able to remove 100-nt introns from pre-mRNA, analogous to the spliceosome. These spliceozymes were improved in this study by 10 rounds of evolution in Escherichia coli cells. One clone with increased activity in E. coli cells was analyzed in detail. Three of its 10 necessary mutations extended the substrate binding duplexes, which led to increased product formation and reduced cleavage at the 5′-splice site. One mutation in the conserved core of the spliceozyme led to a further reduction of cleavage at the 5′-splice site but an increase in cleavage side products at the 3′-splice site. The latter was partially reduced by six additional mutations. Together, the mutations increased product formation while reducing activity at the 5′-splice site and increasing activity at the 3′-splice site. These results show the adaptation of a ribozyme that evolved in nature for cis-splicing to trans-splicing, and they highlight the interdependent function of nucleotides within group I intron ribozymes. Implications for the possible use of spliceozymes as tools in research and therapy, and as a model for the evolution of the spliceosome, are discussed.  相似文献   

14.
Eukaryotic mRNA translation begins with recruitment of the 40S ribosome complex to the mRNA 5′ end through the eIF4F initiation complex binding to the 5′ m7G-mRNA cap. Spliced leader (SL) RNA trans splicing adds a trimethylguanosine (TMG) cap and a sequence, the SL, to the 5′ end of mRNAs. Efficient translation of TMG-capped mRNAs in nematodes requires the SL sequence. Here we define a core set of nucleotides and a stem-loop within the 22-nucleotide nematode SL that stimulate translation of mRNAs with a TMG cap. The structure and core nucleotides are conserved in other nematode SLs and correspond to regions of SL1 required for early Caenorhabditis elegans development. These SL elements do not facilitate translation of m7G-capped RNAs in nematodes or TMG-capped mRNAs in mammalian or plant translation systems. Similar stem-loop structures in phylogenetically diverse SLs are predicted. We show that the nematode eukaryotic translation initiation factor 4E/G (eIF4E/G) complex enables efficient translation of the TMG-SL RNAs in diverse in vitro translation systems. TMG-capped mRNA translation is determined by eIF4E/G interaction with the cap and the SL RNA, although the SL does not increase the affinity of eIF4E/G for capped RNA. These results suggest that the mRNA 5′ untranslated region (UTR) can play a positive and novel role in translation initiation through interaction with the eIF4E/G complex in nematodes and raise the issue of whether eIF4E/G-RNA interactions play a role in the translation of other eukaryotic mRNAs.Cap-dependent translation initiation in eukaryotes is a complex process involving many factors and serves as the primary mechanism for eukaryotic translation (37, 44). The first step in the initiation process, recruitment of the m7G (7-methylguanosine)-capped mRNA to the ribosome, is widely considered the rate-limiting step. It begins with recognition of and binding to the m7G cap at the 5′ end of the mRNA by the eukaryotic translation initiation factor 4F (eIF4F) complex, which contains three proteins: eIF4E (a cap-binding protein), eIF4G (a scaffold protein with RNA binding sites), and eIF4A (an RNA helicase). eIF4G''s interaction with eIF3, itself a multisubunit complex that interacts with the 40S ribosome, facilitates the actual recruitment of capped RNA to the ribosome. With the help of several other initiation factors, the small ribosomal subunit scans the mRNA from 5′ to 3′ until a translation initiation codon (AUG) in appropriate context is identified and an 80S ribosomal complex is formed, after which the first peptide bond is formed, thus ending the initiation process (37, 44). The AUG context can play an important role in the efficiency of translation initiation (23, 44). The length, structure, and presence of AUGs or open reading frames in the mRNA 5′ untranslated region (UTR) can negatively affect cap-dependent translation and ribosomal scanning. In general, long and highly structured 5′ UTRs, as well as upstream AUGs leading to short open reading frames, can impede ribosome scanning and lead to reduced translation (23, 44). In addition, 5′ UTRs less than 10 nucleotides (nt) in length are thought to be too short to enable preinitiation complex assembly and scanning (24). Thus, several attributes of the mRNA 5′ UTR are known to negatively affect translation initiation, whereas only the AUG context and the absence of negative elements are known to have a positive effect on translation initiation (44).Two of the important mRNA features associated with cap-dependent translation, the cap and the 5′ UTR, are significantly altered by an RNA processing event known as spliced leader (SL) trans splicing (3, 8, 17, 26, 36, 47). This takes place in members of a diverse group of eukaryotic organisms, including some protozoa, sponges, cnidarians, chaetognaths, flatworms, nematodes, rotifers, crustaceans, and tunicates (17, 28, 39, 55, 56). In SL trans splicing, a separately transcribed small exon (16 to 51 nucleotides [nt]) with its own cap gets added to the 5′ end of pre-mRNAs. This produces mature mRNAs with a unique cap and a conserved sequence in the 5′ UTR. In metazoa, the m7G cap is replaced with a trimethylguanosine (TMG) cap (m2,2,7GpppN) (27, 30, 46, 49). In nematodes, ∼70% of all mRNAs are trans spliced and therefore have a TMG cap and an SL (2). In general, eukaryotic eIF4E proteins do not effectively recognize the TMG cap (35). This raises the issues of how the translation machinery in trans-splicing metazoa effectively recognizes TMG-capped trans-spliced mRNAs, what role the SL sequence plays in translation initiation, and how the conserved translation initiation machinery has adapted to effectively translate trans-spliced mRNAs.Previous work has shown that efficient translation of TMG-capped messages in nematodes requires the SL sequence (22 nt) immediately downstream of the cap (5, 25, 29). In the current studies, we sought to understand the manner in which the SL enhanced the translation of TMG-capped mRNAs. Using a cell-free nematode in vitro translation system, we carried out mutational analyses that define the specific sequences in the SL that are required and sufficient for efficient translation of TMG-capped mRNAs. These analyses led to the discovery of a small, discrete stem-loop immediately adjacent to the TMG cap in trans-spliced messages required for efficient translation. Notably, the sequences involved in the base pairing of the stem are highly conserved in alternative SL sequences found in nematodes. We further show that the nematode eIF4E/G complex plays a major role in facilitating the SL enhancement of TMG-capped mRNA that likely occurs after the initial cap-binding step. The results demonstrate the importance of specific enhancing elements in the 5′ UTR and adaptation in the eIF4F complex necessary for optimal cap-dependent translation.  相似文献   

15.
16.
17.
Sierro N  Li S  Suzuki Y  Yamashita R  Nakai K 《Gene》2009,430(1-2):44-49
Ciona intestinalis is a useful model organism to analyze chordate development and genetics. However, unlike vertebrates, it shares a unique mechanism called trans-splicing with lower eukaryotes. In the computational analysis of trans-splicing in C. intestinalis we report here, we discovered that although the amount of non-trans-spliced and trans-spliced genes is usually equivalent, the expression ratio between the two groups varies significantly with tissues and developmental stages. Among the seven tissues studied, the observed ratios ranged from 2.53 in “gonad” to 19.53 in “endostyle”, and during development they increased from 1.68 at the “egg” stage to 7.55 at the “juvenile” stage. We further hypothesize that this enrichment in trans-spliced mRNAs in early developmental stages might be related to the abundance of trans-spliced mRNAs in “gonad”. Our analysis indicates that in C. intestinalis, although there may not exist strong fundamental requirements for genes to be trans-spliced, the populations of non-trans-spliced and trans-spliced genes are likely to be spatially and temporally regulated differently.  相似文献   

18.
19.
20.
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