Monday, March 1, 2010

mediated RNA trans-splicing

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 (ΔF508) 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.

Regulation of transcription of the RNA splicing factor hSlu7 by Elk-1 and Sp1 affects alternative splicing

Alternative splicing plays a major role in transcriptome diversity and plasticity, but it is largely unknown how tissue-specific and embryogenesis-specific alternative splicing is regulated. The highly conserved splicing factor Slu7 is involved in 3′ splice site selection and also regulates alternative splicing. We show that Slu7 has a unique spatial pattern of expression among human and mouse embryonic and adult tissues. We identified several functional Ets binding sites and GC-boxes in the human Slu7 (hSlu7) promoter region. The Ets and GC-box binding transcription factors, Elk-1 and Sp1, respectively, exerted opposite effects on hSlu7 transcription: Sp1 protein enhances and Elk-1 protein represses transcription in a dose-dependent manner. Sp1 protein bound to the hSlu7 promoter in vivo, and depletion of Sp1 by RNA interference (RNAi) repressed hSlu7 expression. Elk-1 protein bound to the hSlu7 promoter in vivo, and depletion of Elk-1 by RNAi caused an increase in the endogenous level of hSlu7 mRNA. Further, depletion of either Sp1 or Elk-1 affected alternative splicing. Our results provide indications of a complex transcription regulation mechanism that controls the spatial and temporal expression of Slu7, presumably allowing regulation of tissue-specific alternative splicing events.

Flexibility in the site of exon junction complex deposition revealed by functional group and RNA secondary structure alterations in the splicing subst

The exon junction complex (EJC) is critical for mammalian nonsense-mediated mRNA decay and translational regulation, but the mechanism of its stable deposition on mRNA is unknown. To examine requirements for EJC deposition, we created splicing substrates containing either DNA nucleotides or RNA secondary structure in the 5′ exon. Using RNase H protection, toeprinting, and coimmunoprecipitation assays, we found that EJC location shifts upstream when a stretch of DNA or RNA secondary structure appears at the canonical deposition site. These upstream shifts occur prior to exon ligation and are often accompanied by decreases in deposition efficiency. Although the EJC core protein eIF4AIII contacts four ribose 2′OH groups in crystal structures, we demonstrate that three 2′OH groups are sufficient for deposition. Thus, the site of EJC deposition is more flexible than previously appreciated and efficient deposition appears spatially limited.

Alternately spliced WT1 antisense transcripts interact with WT1 sense RNA and show epigenetic and splicing defects in cancer

Many mammalian genes contain overlapping antisense RNAs, but the functions and mechanisms of action of these transcripts are mostly unknown. WT1 is a well-characterized developmental gene that is mutated in Wilms’ tumor (WT) and acute myeloid leukaemia (AML) and has an antisense transcript (WT1-AS), which we have previously found to regulate WT1 protein levels. In this study, we show that WT1-AS is present in multiple spliceoforms that are usually expressed in parallel with WT1 RNA in human and mouse tissues. We demonstrate that the expression of WT1-AS correlates with methylation of the antisense regulatory region (ARR) in WT1 intron 1, displaying imprinted monoallelic expression in normal kidney and loss of imprinting in WT. However, we find no evidence for imprinting of mouse Wt1-as. WT1-AS transcripts are exported into the cytoplasm and form heteroduplexes with WT1 mRNA in the overlapping region in WT1 exon 1. In AML, there is often abnormal splicing of WT1-AS, which may play a role in the development of this malignancy. These results show that WT1 encodes conserved antisense RNAs that may have an important regulatory role in WT1 expression via RNA:RNA interactions, and which can become deregulated by a variety of mechanisms in cancer.

Analysis of the requirement for RNA polymerase II CTD heptapeptide repeats in pre-mRNA splicing and 3′-end cleavage

The carboxyl-terminal domain (CTD) of RNA polymerase II (pol II) plays an important role in coupling transcription with precursor messenger RNA (pre-mRNA) processing. Efficient capping, splicing, and 3′-end cleavage of pre-mRNA depend on the CTD. Moreover, specific processing factors are known to associate with this structure. The CTD is therefore thought to act as a platform that facilitates the assembly of complexes required for the processing of nascent transcripts. The mammalian CTD contains 52 tandemly repeated heptapeptides with the consensus sequence YSPTSPS. The C-terminal half of the mammalian CTD contains mostly repeats that diverge from this consensus sequence, whereas the N-terminal half contains mostly repeats that match the consensus sequence. Here, we demonstrate that 22 tandem repeats, from either the conserved or divergent halves of the CTD, are sufficient for approximate wild-type levels of transcription, splicing, and 3′-end cleavage of two different pre-mRNAs, one containing a constitutively spliced intron, and the other containing an intron that depends on an exon enhancer for efficient splicing. In contrast, each block of 22 repeats is not sufficient for efficient inclusion of an alternatively spliced exon in another pre-mRNA. In this case, a longer CTD is important for counteracting the negative effect of a splicing silencer element located within the alternative exon. Our results indicate that the length, rather than the composition of CTD repeats, can be the major determinant in efficient processing of different pre-mRNA substrates. However, the extent of this length requirement depends on specific sequence features within the pre-mRNA substrate.

Two reactions of Haloferax volcanii RNA splicing enzymes: Joining of exons and circularization of introns

Archaeal RNA splicing involves at least two protein enzymes, a specific endonuclease and a specific ligase. The endonuclease recognizes and cleaves within a characteristic bulge-helix-bulge (BHB) structure formed by pairing of the regions near the two exon–intron junctions, producing 2‘,3′-cyclic phosphate and 5′-hydroxyl termini. The ligase joins the exons and converts the cyclic phosphate into junction phosphate. The ligated product contains a seven-base hairpin loop, in which the splice junction is in between the two 3′ terminal residues of the loop. Archaeal splicing endonucleases are also involved in rRNA processing, cutting within the BHB structures formed by pairing of the 5′ and 3′ flanking regions of the rRNAs. Large free introns derived from pre-rRNAs have been observed as stable and abundant circular RNAs in certain Crenarchaeota, a kingdom in the domain Archaea. In the present study, we show that the cells of Haloferax volcanii, a Euryarchaeote, contain circular RNAs formed by 3′,5′-phosphodiester linkage between the two termini of the introns derived from their pre-tRNAs. H. volcanii ligase, in vitro, can also circularize both endonuclease-cleaved introns, and non-endonuclease-produced substrates. Exon joining and intron circularization are mechanistically similar ligation reactions that can occur independently. The size of the ligated hairpin loop and position of the splice junction within this loop can be changed in in vitro ligation reactions. Overall, archaeal RNA splicing seems to involve two sets of two symmetric transesterification reactions each.

B-cell and plasma-cell splicing differences: A potential role in regulated immunoglobulin RNA processing

The immunoglobulin μ pre-mRNA is alternatively processed at its 3′ end by competing splice and cleavage-polyadenylation reactions to generate mRNAs encoding the membrane-associated or secreted forms of the IgM protein, respectively. The relative use of the competing processing pathways varies during B-lymphocyte development, and it has been established previously that cleavage-polyadenylation activity is higher in plasma cells, which secrete IgM, than in B cells, which produce membrane-associated IgM. To determine whether RNA-splicing activity varies during B-lymphocyte development to contribute to μ RNA-processing regulation, we first demonstrate that μ pre-mRNA processing is sensitive to artificial changes in the splice environment by coexpressing SR proteins with the μ gene. To explore differences between the splice environments of B cells and plasma cells, we analyzed the splicing patterns from two different chimeric non-Ig genes that can be alternatively spliced but have no competing cleavage-polyadenylation reaction. The ratio of intact exon splicing to cryptic splice site use from one chimeric gene differs between several B-cell and several plasma-cell lines. Also, the amount of spliced RNA is higher in B-cell than plasma-cell lines from a set of genes whose splicing is dependent on a functional exonic splice enhancer. Thus, there is clear difference between the B-cell and plasma-cell splicing environments. We propose that both general cleavage-polyadenylation and general splice activities are modulated during B-lymphocyte development to ensure proper regulation of the alternative μ RNA processing pathways.

Conserved RNA secondary structures promote alternative splicing

Pre-mRNA splicing is carried out by the spliceosome, which identifies exons and removes intervening introns. Alternative splicing in higher eukaryotes results in the generation of multiple protein isoforms from gene transcripts. The extensive alternative splicing observed implies a flexibility of the spliceosome to identify exons within a given pre-mRNA. To reach this flexibility, splice-site selection in higher eukaryotes has evolved to depend on multiple parameters such as splice-site strength, splicing regulators, the exon/intron architecture, and the process of pre-mRNA synthesis itself. RNA secondary structures have also been proposed to influence alternative splicing as stable RNA secondary structures that mask splice sites are expected to interfere with splice-site recognition. Using structural and functional conservation, we identified RNA structure elements within the human genome that associate with alternative splice-site selection. Their frequent involvement with alternative splicing demonstrates that RNA structure formation is an important mechanism regulating gene expression and disease.

Efficient and specific repair of sickle β-globin RNA by trans-splicing ribozymes

Previously we demonstrated that a group I ribozyme can perform trans-splicing to repair sickle β-globin transcripts upon transfection of in vitro transcribed ribozyme into mammalian cells. Here, we sought to develop expression cassettes that would yield high levels of active ribozyme after gene transfer. Our initial expression constructs were designed to generate trans-slicing ribozymes identical to those used in our previous RNA transfection studies with ribozymes containing 6-nucleotide long internal guide sequences. The ribozymes expressed from these cassettes, however, were found to be unable to repair sickle β-globin RNAs. Further experiments revealed that two additional structural elements are important for ribozyme-mediate RNA repair: the P10 interaction formed between the 5′ end of the ribozyme and the beginning of the 3′ exon and an additional base-pairing interaction formed between an extended guide sequence and the substrate RNA. These optimized expression cassettes yield ribozymes that are able to amend 10%–50% of the sickle β-globin RNAs in transfected mammalian cells. Finally, a ribozyme with a 5-bp extended guide sequence preferentially reacts with sickle β-globin RNAs over wild-type β-globin RNAs, although the wild-type β-globin transcript forms only a single mismatch with the ribozyme. These results demonstrate that trans-splicing ribozyme expression cassettes can be generated to yield ribozymes that can repair a clinically relevant fraction of sickle β-globin RNAs in mammalian cells with greatly improved specificity.

Polyadenylation releases mRNA from RNA polymerase II in a process that is licensed by splicing

When transcription is coupled to pre-mRNA processing in HeLa nuclear extracts nascent transcripts become attached to RNA polymerase II during assembly of the cleavage/polyadenylation apparatus (CPA), and are not released even after cleavage at the poly(A) site. Here we show that these cleaved transcripts are anchored to the polymerase at their 3′ ends by the CPA or, when introns are present, by the larger 3′-terminal exon definition complex (EDC), which consists of splicing factors complexed with the CPA. Poly(A) addition releases the RNA from the polymerase when the RNA is anchored only by the CPA. When anchored by the EDC, poly(A) addition remains a requirement, but it triggers release only after being licensed by splicing. The process by which RNA must first be attached to the polymerase by the EDC, and then can only be released following dual inputs from splicing and polyadenylation, provides an obvious opportunity for surveillance as the RNA enters the transport pathway.

Genetic identification of potential RNA-binding regions in a group II intron-encoded reverse transcriptase

Mobile group II introns encode a reverse transcriptase that binds the intron RNA to promote RNA splicing and intron mobility, the latter via reverse splicing of the excised intron into DNA sites, followed by reverse transcription. Previous work showed that the Lactococcus lactis Ll.LtrB intron reverse transcriptase, denoted LtrA protein, binds with high affinity to DIVa, a stem–loop structure at the beginning of the LtrA open reading frame and makes additional contacts with intron core regions that stabilize the active RNA structure for forward and reverse splicing. LtrA's binding to DIVa down-regulates its translation and is critical for initiation of reverse transcription. Here, by using high-throughput unigenic evolution analysis with a genetic assay in which LtrA binding to DIVa down-regulates translation of GFP, we identified regions at LtrA's N terminus that are required for DIVa binding. Then, by similar analysis with a reciprocal genetic assay, we confirmed that residual splicing of a mutant intron lacking DIVa does not require these N-terminal regions, but does require other reverse transcriptase (RT) and X/thumb domain regions that bind the intron core. We also show that N-terminal fragments of LtrA by themselves bind specifically to DIVa in vivo and in vitro. Our results suggest a model in which the N terminus of nascent LtrA binds DIVa of the intron RNA that encoded it and nucleates further interactions with core regions that promote RNP assembly for RNA splicing and intron mobility. Features of this model may be relevant to evolutionarily related non-long-terminal-repeat (non-LTR)-retrotransposon RTs.