Alternative splicing

Alternative splicing produces three protein isoforms.

Alternative splicing, or differential splicing, is a regulated process during gene expression that results in a single gene coding for multiple proteins. In this process, particular exons of a gene may be included within or excluded from the final, processed messenger RNA (mRNA) produced from that gene.[1] Consequently, the proteins translated from alternatively spliced mRNAs will contain differences in their amino acid sequence and, often, in their biological functions (see Figure). Notably, alternative splicing allows the human genome to direct the synthesis of many more proteins than would be expected from its 20,000 protein-coding genes.

Alternative splicing occurs as a normal phenomenon in eukaryotes, where it greatly increases the biodiversity of proteins that can be encoded by the genome;[1] in humans, ~95% of multi-exonic genes are alternatively spliced.[2] There are numerous modes of alternative splicing observed, of which the most common is exon skipping. In this mode, a particular exon may be included in mRNAs under some conditions or in particular tissues, and omitted from the mRNA in others.[1]

The production of alternatively spliced mRNAs is regulated by a system of trans-acting proteins that bind to cis-acting sites on the primary transcript itself. Such proteins include splicing activators that promote the usage of a particular splice site, and splicing repressors that reduce the usage of a particular site. Mechanisms of alternative splicing are highly variable, and new examples are constantly being found, particularly through the use of high-throughput techniques. Researchers hope to fully elucidate the regulatory systems involved in splicing, so that alternative splicing products from a given gene under particular conditions ("splicing variants") could be predicted by a "splicing code".[3][4]

Abnormal variations in splicing are also implicated in disease; a large proportion of human genetic disorders result from splicing variants.[3] Abnormal splicing variants are also thought to contribute to the development of cancer,[5][6][7][8] and splicing factor genes are frequently mutated in different types of cancer.[8]

Discovery

Alternative splicing was first observed in 1977.[9][10] The Adenovirus produces five primary transcripts early in its infectious cycle, prior to viral DNA replication, and an additional one later, after DNA replication begins. The early primary transcripts continue to be produced after DNA replication begins. The additional primary transcript produced late in infection is large and comes from 5/6 of the 32kb adenovirus genome. This is much larger than any of the individual adenovirus mRNAs present in infected cells. Researchers found that the primary RNA transcript produced by adenovirus type 2 in the late phase was spliced in many different ways, resulting in mRNAs encoding different viral proteins. In addition, the primary transcript contained multiple polyadenylation sites, giving different 3’ ends for the processed mRNAs.[11][12][13]

In 1981, the first example of alternative splicing in a transcript from a normal, endogenous gene was characterized.[11] The gene encoding the thyroid hormone calcitonin was found to be alternatively spliced in mammalian cells. The primary transcript from this gene contains 6 exons; the calcitonin mRNA contains exons 1–4, and terminates after a polyadenylation site in exon 4. Another mRNA is produced from this pre-mRNA by skipping exon 4, and includes exons 1–3, 5, and 6. It encodes a protein known as CGRP (calcitonin gene related peptide).[14][15] Examples of alternative splicing in immunoglobin gene transcripts in mammals were also observed in the early 1980s.[11][16]

Since then, alternative splicing has been found to be ubiquitous in eukaryotes.[1] The "record-holder" for alternative splicing is a D. melanogaster gene called Dscam, which could potentially have 38,016 splice variants.[17]

Modes

Traditional classification of basic types of alternative RNA splicing events.
Relative frequencies of types of alternative splicing events differ between humans and fruit flies.[18]

Five basic modes of alternative splicing are generally recognized.[1][2][3][18]

In addition to these primary modes of alternative splicing, there are two other main mechanisms by which different mRNAs may be generated from the same gene; multiple promoters and multiple polyadenylation sites. Use of multiple promoters is properly described as a transcriptional regulation mechanism rather than alternative splicing; by starting transcription at different points, transcripts with different 5'-most exons can be generated. At the other end, multiple polyadenylation sites provide different 3' end points for the transcript. Both of these mechanisms are found in combination with alternative splicing and provide additional variety in mRNAs derived from a gene.[1][3]

Schematic cutoff from 3 splicing structures in the murine hyaluronidase gene. Directionality of transcription from 5' to 3' is shown from left to right. Exons and introns are not drawn to scale.


These modes describe basic splicing mechanisms, but may be inadequate to describe complex splicing events. For instance, the figure to the right shows 3 spliceforms from the mouse hyaluronidase 3 gene. Comparing the exonic structure shown in the first line (green) with the one in the second line (yellow) shows intron retention, whereas the comparison between the second and the third spliceform (yellow vs. blue) exhibits exon skipping. A model nomenclature to uniquely designate all possible splicing patterns has recently been proposed.[18]

Alternative splicing mechanisms

General splicing mechanism

Main article: RNA splicing
Spliceosome A complex defines the 5' and 3' ends of the intron before removal[3]

When the pre-mRNA has been transcribed from the DNA, it includes several introns and exons. (In nematodes, the mean is 4–5 exons and introns; in the fruit fly Drosophila there can be more than 100 introns and exons in one transcribed pre-mRNA.) The exons to be retained in the mRNA are determined during the splicing process. The regulation and selection of splice sites are done by trans-acting splicing activator and splicing repressor proteins as well as cis-acting elements within the pre-mRNA itself such as exonic splicing enhancers and exonic splicing silencers.

The typical eukaryotic nuclear intron has consensus sequences defining important regions. Each intron has GU at its 5' end. Near the 3' end there is a branch site. The nucleotide at the branchpoint is always an A; the consensus around this sequence varies somewhat. In humans the branch site consensus sequence is yUnAy.[19] The branch site is followed by a series of pyrimidines - the polypyrimidine tract - then by AG at the 3' end.[3]

Splicing of mRNA is performed by an RNA and protein complex known as the spliceosome, containing snRNPs designated U1, U2, U4, U5, and U6 (U3 is not involved in mRNA splicing).[20] U1 binds to the 5' GU and U2, with the assistance of the U2AF protein factors, binds to the branchpoint A within the branch site. The complex at this stage is known as the spliceosome A complex. Formation of the A complex is usually the key step in determining the ends of the intron to be spliced out, and defining the ends of the exon to be retained.[3] (The U nomenclature derives from their high uridine content).

The U4,U5,U6 complex binds, and U6 replaces the U1 position. U1 and U4 leave. The remaining complex then performs two transesterification reactions. In the first transesterification, 5' end of the intron is cleaved from the upstream exon and joined to the branch site A by a 2',5'-phosphodiester linkage. In the second transesterification, the 3' end of the intron is cleaved from the downstream exon, and the two exons are joined by a phosphodiester bond. The intron is then released in lariat form and degraded.[1]

Regulatory elements and proteins

Splicing repression

Splicing is regulated by trans-acting proteins (repressors and activators) and corresponding cis-acting regulatory sites (silencers and enhancers) on the pre-mRNA. However, as part of the complexity of alternative splicing, it is noted that the effects of a splicing factor are frequently position-dependent. That is, a splicing factor that serves as a splicing activator when bound to an intronic enhancer element may serve as a repressor when bound to its splicing element in the context of an exon, and vice versa.[21] The secondary structure of the pre-mRNA transcript also plays a role in regulating splicing, such as by bringing together splicing elements or by masking a sequence that would otherwise serve as a binding element for a splicing factor.[22][23] Together, these elements form a "splicing code" that governs how splicing will occur under different cellular conditions.[24][25]

There are two major types of cis-acting RNA sequence elements present in pre-mRNAs and they have corresponding trans-acting RNA-binding proteins. Splicing silencers are sites to which splicing repressor proteins bind, reducing the probability that a nearby site will be used as a splice junction. These can be located in the intron itself (intronic splicing silencers, ISS) or in a neighboring exon (exonic splicing silencers, ESS). They vary in sequence, as well as in the types of proteins that bind to them. The majority of splicing repressors are heterogeneous nuclear ribonucleoproteins (hnRNPs) such as hnRNPA1 and polypyrimidine tract binding protein (PTB).[3][24] Splicing enhancers are sites to which splicing activator proteins bind, increasing the probability that a nearby site will be used as a splice junction. These also may occur in the intron (intronic splicing enhancers, ISE) or exon (exonic splicing enhancers, ESE). Most of the activator proteins that bind to ISEs and ESEs are members of the SR protein family. Such proteins contain RNA recognition motifs and arginine and serine-rich (RS) domains.[3][24]

Splicing activation

In general, the determinants of splicing work in an inter-dependent manner that depends on context, so that the rules governing how splicing is regulated from a splicing code.[25] The presence of a particular cis-acting RNA sequence element may increase the probability that a nearby site will be spliced in some cases, but decrease the probability in other cases, depending on context. The context within which regulatory elements act includes cis-acting context that is established by the presence of other RNA sequence features, and trans-acting context that is established by cellular conditions. For example, some cis-acting RNA sequence elements influence splicing only if multiple elements are present in the same region so as to establish context. As another example, a cis-acting element can have opposite effects on splicing, depending on which proteins are expressed in the cell (e.g., neuronal versus non-neuronal PTB). The adaptive significance of splicing silencers and enhancers is attested by studies showing that there is strong selection in human genes against mutations that produce new silencers or disrupt existing enhancers.[26][27]

Examples

Exon skipping: Drosophila dsx

Alternative splicing of dsx pre-mRNA

Pre-mRNAs from the D. melanogaster gene dsx contain 6 exons. In males, exons 1,2,3,5,and 6 are joined to form the mRNA, which encodes a transcriptional regulatory protein required for male development. In females, exons 1,2,3, and 4 are joined, and a polyadenylation signal in exon 4 causes cleavage of the mRNA at that point. The resulting mRNA is a transcriptional regulatory protein required for female development.[28]

This is an example of exon skipping. The intron upstream from exon 4 has a polypyrimidine tract that doesn't match the consensus sequence well, so that U2AF proteins bind poorly to it without assistance from splicing activators. This 3' splice acceptor site is therefore not used in males. Females, however, produce the splicing activator Transformer (Tra) (see below). The SR protein Tra2 is produced in both sexes and binds to an ESE in exon 4; if Tra is present, it binds to Tra2 and, along with another SR protein, forms a complex that assists U2AF proteins in binding to the weak polypyrimidine tract. U2 is recruited to the associated branchpoint, and this leads to inclusion of exon 4 in the mRNA.[28][29]

Alternative acceptor sites: Drosophila Transformer

Alternative splicing of the Drosophila Transformer gene product.

Pre-mRNAs of the Transformer (Tra) gene of Drosophila melanogaster undergo alternative splicing via the alternative acceptor site mode. The gene Tra encodes a protein that is expressed only in females. The primary transcript of this gene contains an intron with two possible acceptor sites. In males, the upstream acceptor site is used. This causes a longer version of exon 2 to be included in the processed transcript, including an early stop codon. The resulting mRNA encodes a truncated protein product that is inactive. Females produce the master sex determination protein Sex lethal (Sxl). The Sxl protein is a splicing repressor that binds to an ISS in the RNA of the Tra transcript near the upstream acceptor site, preventing U2AF protein from binding to the polypyrimidine tract. This prevents the use of this junction, shifting the spliceosome binding to the downstream acceptor site. Splicing at this point bypasses the stop codon, which is excised as part of the intron. The resulting mRNA encodes an active Tra protein, which itself is a regulator of alternative splicing of other sex-related genes (see dsx above).[1]

Exon definition: Fas receptor

Alternative splicing of the Fas receptor pre-mRNA

Multiple isoforms of the Fas receptor protein are produced by alternative splicing. Two normally occurring isoforms in humans are produced by an exon-skipping mechanism. An mRNA including exon 6 encodes the membrane-bound form of the Fas receptor, which promotes apoptosis, or programmed cell death. Increased expression of Fas receptor in skin cells chronically exposed to the sun, and absence of expression in skin cancer cells, suggests that this mechanism may be important in elimination of pre-cancerous cells in humans.[30] If exon 6 is skipped, the resulting mRNA encodes a soluble Fas protein that does not promote apoptosis. The inclusion or skipping of the exon depends on two antagonistic proteins, TIA-1 and polypyrimidine tract-binding protein (PTB).

This mechanism is an example of exon definition in splicing. A spliceosome assembles on an intron, and the snRNP subunits fold the RNA so that the 5' and 3' ends of the intron are joined. However, recently studied examples such as this one show that there are also interactions between the ends of the exon. In this particular case, these exon definition interactions are necessary to allow the binding of core splicing factors prior to assembly of the spliceosomes on the two flanking introns.[31]

Repressor-activator competition: HIV-1 tat exon 2

Alternative splicing of HIV-1 tat exon 2

HIV, the retrovirus that causes AIDS in humans, produces a single primary RNA transcript, which is alternatively spliced in multiple ways to produce over 40 different mRNAs.[32] Equilibrium among differentially spliced transcripts provides multiple mRNAs encoding different products that are required for viral multiplication.[33] One of the differentially spliced transcripts contains the tat gene, in which exon 2 is a cassette exon that may be skipped or included. The inclusion of tat exon 2 in the RNA is regulated by competition between the splicing repressor hnRNP A1 and the SR protein SC35. Within exon 2 an exonic splicing silencer sequence (ESS) and an exonic splicing enhancer sequence (ESE) overlap. If A1 repressor protein binds to the ESS, it initiates cooperative binding of multiple A1 molecules, extending into the 5’ donor site upstream of exon 2 and preventing the binding of the core splicing factor U2AF35 to the polypyrimidine tract. If SC35 binds to the ESE, it prevents A1 binding and maintains the 5’ donor site in an accessible state for assembly of the spliceosome. Competition between the activator and repressor ensures that both mRNA types (with and without exon 2) are produced.[32]

Adaptive significance

Alternative splicing is one of several exceptions to the original idea that one DNA sequence codes for one polypeptide (the One gene-one enzyme hypothesis). It might be more correct now to say "One gene – many polypeptides".[34] External information is needed in order to decide which polypeptide is produced, given a DNA sequence and pre-mRNA. Since the methods of regulation are inherited, this provides novel ways for mutations to affect gene expression.[7]

It has been proposed that for eukaryotes alternative splicing was a very important step towards higher efficiency, because information can be stored much more economically. Several proteins can be encoded by a single gene, rather than requiring a separate gene for each, and thus allowing a more varied proteome from a genome of limited size.[1] It also provides evolutionary flexibility. A single point mutation may cause a given exon to be occasionally excluded or included from a transcript during splicing, allowing production of a new protein isoform without loss of the original protein.[1] Studies have identified intrinsically disordered regions (see Intrinsically unstructured proteins) as enriched in the non-constitutive exons[35] suggesting that protein isoforms may display functional diversity due to the alteration of functional modules within these regions. Such functional diversity achieved by isoforms is reflected by their expression patterns and can be predicted by machine learning approaches.[36][37] Comparative studies indicate that alternative splicing preceded multicellularity in evolution, and suggest that this mechanism might have been co-opted to assist in the development of multicellular organisms.[38]

Research based on the Human Genome Project and other genome sequencing has shown that humans have only about 30% more genes than the roundworm Caenorhabditis elegans, and only about twice as many as the fly Drosophila melanogaster. This finding led to speculation that the perceived greater complexity of humans, or vertebrates generally, might be due to higher rates of alternative splicing in humans than are found in invertebrates.[39][40] However, a study on samples of 100,000 ESTs each from human, mouse, rat, cow, fly (D. melanogaster), worm (C. elegans), and the plant Arabidopsis thaliana found no large differences in frequency of alternatively spliced genes among humans and any of the other animals tested.[41] Another study, however, proposed that these results were an artifact of the different numbers of ESTs available for the various organisms. When they compared alternative splicing frequencies in random subsets of genes from each organism, the authors concluded that vertebrates do have higher rates of alternative splicing than invertebrates.[42]

Alternative splicing and disease

Changes in the RNA processing machinery may lead to mis-splicing of multiple transcripts, while single-nucleotide alterations in splice sites or cis-acting splicing regulatory sites may lead to differences in splicing of a single gene, and thus in the mRNA produced from a mutant gene's transcripts. A study in 2005 involving probabilistic analyses indicated that greater than 60% of human disease-causing mutations affect splicing rather than directly affecting coding sequences.[43] A more recent study indicates that one-third of all hereditary diseases are likely to have a splicing component.[21] Regardless of exact percentage, a number of splicing-related diseases do exist.[44] As described below, a prominent example of splicing-related diseases is cancer.

Abnormally spliced mRNAs are also found in a high proportion of cancerous cells.[5][6][8] Combined RNA-Seq and proteomics analyses have revealed striking differential expression of splice isoforms of key proteins in important cancer pathways.[45] It is not always clear whether such aberrant patterns of splicing contribute to the cancerous growth, or are merely consequence of cellular abnormalities associated with cancer. Interestingly, for certain types of cancer, like in colorectal and prostate, the number of splicing errors per cancer has been shown to vary greatly between individual cancers, a phenomenon referred to as transcriptome instability.[46][47] Transcriptome instability has further been shown to correlate grealty with reduced expression level of splicing factor genes. In fact, there is actually a reduction of alternative splicing in cancerous cells compared to normal ones, and the types of splicing differ; for instance, cancerous cells show higher levels of intron retention than normal cells, but lower levels of exon skipping.[48] Some of the differences in splicing in cancerous cells may be due to the high frequency of somatic mutations in splicing factor genes,[8] and some may result from changes in phosphorylation of trans-acting splicing factors.[7] Others may be produced by changes in the relative amounts of splicing factors produced; for instance, breast cancer cells have been shown to have increased levels of the splicing factor SF2/ASF.[49] One study found that a relatively small percentage (383 out of over 26000) of alternative splicing variants were significantly higher in frequency in tumor cells than normal cells, suggesting that there is a limited set of genes which, when mis-spliced, contribute to tumor development.[50] It is believed however that the deleterious effects of mis-spliced transcripts are usually safeguarded and eliminated by a cellular posttranscriptional quality control mechanism termed Nonsense-mediated mRNA decay [NMD].[51]

One example of a specific splicing variant associated with cancers is in one of the human DNMT genes. Three DNMT genes encode enzymes that add methyl groups to DNA, a modification that often has regulatory effects. Several abnormally spliced DNMT3B mRNAs are found in tumors and cancer cell lines. In two separate studies, expression of two of these abnormally spliced mRNAs in mammalian cells caused changes in the DNA methylation patterns in those cells. Cells with one of the abnormal mRNAs also grew twice as fast as control cells, indicating a direct contribution to tumor development by this product.[7]

Another example is the Ron (MST1R) proto-oncogene. An important property of cancerous cells is their ability to move and invade normal tissue. Production of an abnormally spliced transcript of Ron has been found to be associated with increased levels of the SF2/ASF in breast cancer cells. The abnormal isoform of the Ron protein encoded by this mRNA leads to cell motility.[49]

Overexpression of a truncated splice variant of the FOSB gene – ΔFosB – in a specific population of neurons in the nucleus accumbens has been identified as the causal mechanism involved in the induction and maintenance of an addiction to drugs and natural rewards.[52][53][54][55]

Recent provocative studies point to a key function of chromatin structure and histone modifications in alternative splicing regulation. These insights suggest that epigenetic regulation determines not only what parts of the genome are expressed but also how they are spliced.[56]

Genome-wide analysis of alternative splicing

Genome-wide analysis of alternative splicing is a challenging task. Typically, alternatively spliced transcripts have been found by comparing EST sequences, but this requires sequencing of very large numbers of ESTs. Most EST libraries come from a very limited number of tissues, so tissue-specific splice variants are likely to be missed in any case. High-throughput approaches to investigate splicing have, however, been developed, such as: DNA microarray-based analyses, RNA-binding assays, and deep sequencing. These methods can be used to screen for polymorphisms or mutations in or around splicing elements that affect protein binding. When combined with splicing assays, including in vivo reporter gene assays, the functional effects of polymorphisms or mutations on the splicing of pre-mRNA transcripts can then be analyzed.[21][24][57]

In microarray analysis, arrays of DNA fragments representing individual exons (e.g. Affymetrix exon microarray) or exon/exon boundaries (e.g. arrays from ExonHit or Jivan) have been used. The array is then probed with labeled cDNA from tissues of interest. The probe cDNAs bind to DNA from the exons that are included in mRNAs in their tissue of origin, or to DNA from the boundary where two exons have been joined. This can reveal the presence of particular alternatively spliced mRNAs.[58]

CLIP (Cross-linking and immunoprecipitation) uses UV radiation to link proteins to RNA molecules in a tissue during splicing. A trans-acting splicing regulatory protein of interest is then precipitated using specific antibodies. When the RNA attached to that protein is isolated and cloned, it reveals the target sequences for that protein.[4] Another method for identifying RNA-binding proteins and mapping their binding to pre-mRNA transcripts is "Microarray Evaluation of Genomic Aptamers by shift (MEGAshift)".net[59] This method involves an adaptation of the "Systematic Evolution of Ligands by Exponential Enrichment (SELEX)" method[60] together with a microarray-based readout. Use of the MEGAshift method has provided insights into the regulation of alternative splicing by allowing for the identification of sequences in pre-mRNA transcripts surrounding alternatively spliced exons that mediate binding to different splicing factors, such as ASF/SF2 and PTB.[61] This approach has also been used to aid in determining the relationship between RNA secondary structure and the binding of splicing factors.[23]

Deep sequencing technologies have been used to conduct genome-wide analyses of mRNAs - unprocessed and processed - thus providing insights into alternative splicing. For example, results from use of deep sequencing indicate that, in humans, an estimated 95% of transcripts from multiexon genes undergo alternative splicing, with a number of pre-mRNA transcripts spliced in a tissue-specific manner.[2] Functional genomics and computational approaches based on multiple instance learning have also been developed to integrate RNA-seq data to predict functions for alternatively spliced isoforms.[37] Deep sequencing has also aided in the in vivo detection of the transient lariats that are released during splicing, the determination of branch site sequences, and the large-scale mapping of branchpoints in human pre-mRNA transcripts.[62]

Use of reporter assays makes it possible to find the splicing proteins involved in a specific alternative splicing event by constructing reporter genes that will express one of two different fluorescent proteins depending on the splicing reaction that occurs. This method has been used to isolate mutants affecting splicing and thus to identify novel splicing regulatory proteins inactivated in those mutants.[4]

Alternative splicing databases

There is a collection of alternative splicing databases. These databases are useful for finding genes having pre-mRNAs undergoing alternative splicing and alternative splicing events. Such as, the Plant Alternative Splicing Database contains alternative splicing information for several major cereal plant species and other plant species.

See also

References

  1. 1 2 3 4 5 6 7 8 9 10 Black, Douglas L. (2003). "Mechanisms of alternative pre-messenger RNA splicing". Annual Review of Biochemistry. 72 (1): 291–336. doi:10.1146/annurev.biochem.72.121801.161720. PMID 12626338.
  2. 1 2 3 Pan, Q; Shai O; Lee LJ; Frey BJ; Blencowe BJ (Dec 2008). "Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing". Nature Genetics. 40 (12): 1413–1415. doi:10.1038/ng.259. PMID 18978789.
  3. 1 2 3 4 5 6 7 8 9 10 Matlin, AJ; Clark, F; Smith, CWJ (May 2005). "Understanding alternative splicing: towards a cellular code". Nature Reviews. 6 (5): 386–398. doi:10.1038/nrm1645. PMID 15956978.
  4. 1 2 3 David, C. J.; Manley, J. L. (2008). "The search for alternative splicing regulators: new approaches offer a path to a splicing code". Genes & Development. 22 (3): 279–85. doi:10.1101/gad.1643108. PMC 2731647Freely accessible. PMID 18245441.
  5. 1 2 Skotheim, R I; Nees, M (2007). "Alternative splicing in cancer: noise, functional, or systematic?". The international journal of biochemistry & cell biology. 39 (7-8): 1432–49. doi:10.1016/j.biocel.2007.02.016. PMID 17416541.
  6. 1 2 Bauer, Joseph Alan; He, Chunjiang; Zhou, Fang; Zuo, Zhixiang; Cheng, Hanhua; Zhou, Rongjia (2009). Bauer, Joseph Alan, ed. "A Global View of Cancer-Specific Transcript Variants by Subtractive Transcriptome-Wide Analysis". PLoS ONE. 4 (3): e4732. Bibcode:2009PLoSO...4.4732H. doi:10.1371/journal.pone.0004732. PMC 2648985Freely accessible. PMID 19266097.
  7. 1 2 3 4 Fackenthal, Jd; Godley, La (2008). "Aberrant RNA splicing and its functional consequences in cancer cells" (Free full text). Disease models & mechanisms. 1 (1): 37–42. doi:10.1242/dmm.000331. PMC 2561970Freely accessible. PMID 19048051.
  8. 1 2 3 4 Sveen, A; Kilpinen, S; Ruusulehto, A; Lothe, RA; Skotheim, RI (2015). "Aberrant RNA splicing in cancer; expression changes and driver mutations of splicing factor genes". Oncogene. doi:10.1038/onc.2015.318. PMID 26300000.
  9. Chow, Louise T.; Gelinas, Richard E.; Broker, Thomas R.; Roberts, Richard J. (1977). "An amazing sequence arrangement at the 5′ ends of adenovirus 2 messenger RNA". Cell. 12 (1): 1–8. doi:10.1016/0092-8674(77)90180-5. PMID 902310.
  10. Berget SM, Moore C, Sharp PA (1977). "Spliced segments at the 5' terminus of adenovirus 2 late mRNA". Proc. Natl. Acad. Sci. U.S.A. 74 (8): 3171–5. Bibcode:1977PNAS...74.3171B. doi:10.1073/pnas.74.8.3171. PMC 431482Freely accessible. PMID 269380.
  11. 1 2 3 Leff SE, Rosenfeld MG, Evans RM (1986). "Complex transcriptional units: diversity in gene expression by alternative RNA processing". Annu. Rev. Biochem. 55 (1): 1091–117. doi:10.1146/annurev.bi.55.070186.005303. PMID 3017190.
  12. Chow LT, Broker TR (1978). "The spliced structures of adenovirus 2 fiber message and the other late mRNAs". Cell. 15 (2): 497–510. doi:10.1016/0092-8674(78)90019-3. PMID 719751.
  13. Nevins JR, Darnell JE (1978). "Steps in the processing of Ad2 mRNA: poly(A)+ nuclear sequences are conserved and poly(A) addition precedes splicing". Cell. 15 (4): 1477–93. doi:10.1016/0092-8674(78)90071-5. PMID 729004.
  14. Rosenfeld MG, Amara SG, Roos BA, Ong ES, Evans RM (1981). "Altered expression of the calcitonin gene associated with RNA polymorphism". Nature. 290 (5801): 63–5. Bibcode:1981Natur.290...63R. doi:10.1038/290063a0. PMID 7207587.
  15. Rosenfeld MG, Lin CR, Amara SG, et al. (1982). "Calcitonin mRNA polymorphism: peptide switching associated with alternative RNA splicing events". Proc. Natl. Acad. Sci. U.S.A. 79 (6): 1717–21. Bibcode:1982PNAS...79.1717R. doi:10.1073/pnas.79.6.1717. PMC 346051Freely accessible. PMID 6952224.
  16. Maki R, Roeder W, Traunecker A, et al. (1981). "The role of DNA rearrangement and alternative RNA processing in the expression of immunoglobulin delta genes". Cell. 24 (2): 353–65. doi:10.1016/0092-8674(81)90325-1. PMID 6786756.
  17. Schmucker D, Clemens JC, Shu H, Worby CA, Xiao J, Muda M, Dixon JE, Zipursky SL (2000). "Drosophila Dscam is an axon guidance receptor exhibiting extraordinary molecular diversity". Cell. 101 (6): 671–84. doi:10.1016/S0092-8674(00)80878-8. PMID 10892653.
  18. 1 2 3 4 5 Michael Sammeth; Sylvain Foissac; Roderic Guigó (2008). Brent, Michael R., ed. "A general definition and nomenclature for alternative splicing events". PLoS Comput. Biol. 4 (8): e1000147. Bibcode:2008PLSCB...4E0147S. doi:10.1371/journal.pcbi.1000147. PMC 2467475Freely accessible. PMID 18688268.
  19. Gao, K.; Masuda, A.; Matsuura, T.; Ohno, K. (2008). "Human branch point consensus sequence is yUnAy". Nucleic Acids Research. 36 (7): 2257–67. doi:10.1093/nar/gkn073. PMC 2367711Freely accessible. PMID 18285363.
  20. Clark, David (2005). Molecular biology. Amsterdam: Elsevier Academic Press. ISBN 0-12-175551-7.
  21. 1 2 3 Lim, KH; Ferraris, L; Filloux, ME; Raphael, BJ; Fairbrother, WG (2011). "Using positional distribution to identify splicing elements and predict pre-mRNA processing defects in human genes". Proc. Natl. Acad. Sci. USA. 108 (27): 11093–11098. doi:10.1073/pnas.1101135108. PMC 3131313Freely accessible. PMID 21685335.
  22. Warf, MB; Berglund, JA (2010). "Role of RNA structure in regulating pre-mRNA splicing". Trends Biochem. Sci. 35 (3): 169–178. doi:10.1016/j.tibs.2009.10.004. PMC 2834840Freely accessible. PMID 19959365.
  23. 1 2 Reid, DC; Chang, BL; Gunderson, SI; Alpert, L; Thompson, WA; Fairbrother, WG (2009). "Next-generation SELEX identifies sequence and structural determinants of splicing factor binding in human pre-mRNA sequence". RNA. 15 (12): 2385–2397. doi:10.1261/rna.1821809. PMC 2779669Freely accessible. PMID 19861426.
  24. 1 2 3 4 Wang, Z; Burge, Cb (2008). "Splicing regulation: from a parts list of regulatory elements to an integrated splicing code" (Free full text). RNA. 14 (5): 802–13. doi:10.1261/rna.876308. PMC 2327353Freely accessible. PMID 18369186.
  25. 1 2 Barash, Y; et al. (2010). "Deciphering the splicing code". Nature. 465 (7294): 53–59. Bibcode:2010Natur.465...53B. doi:10.1038/nature09000. PMID 20445623.
  26. Ke S, Zhang XH, Chasin LA (2008). "Positive selection acting on splicing motifs reflects compensatory evolution". Genome Res. 18 (4): 533–43. doi:10.1101/gr.070268.107. PMC 2279241Freely accessible. PMID 18204002.
  27. Fairbrother, WG; Holste, D; Burge, CB; Sharp, PA (2004). "Single nucleotide polymorphism–based validation of exonic splicing enhancers". PLoS Biol. 2 (9): e268. doi:10.1371/journal.pbio.0020268. PMC 514884Freely accessible. PMID 15340491.
  28. 1 2 Lynch KW, Maniatis T (1996). "Assembly of specific SR protein complexes on distinct regulatory elements of the Drosophila doublesex splicing enhancer". Genes Dev. 10 (16): 2089–101. doi:10.1101/gad.10.16.2089. PMID 8769651.
  29. Graveley BR, Hertel KJ, Maniatis T (2001). "The role of U2AF35 and U2AF65 in enhancer-dependent splicing". RNA. 7 (6): 806–18. doi:10.1017/S1355838201010317. PMC 1370132Freely accessible. PMID 11421359.
  30. Filipowicz, Ewa; Adegboyega, P.; Sanchez, R. L.; Gatalica, Zoran (2002). "Expression of CD95 (Fas) in sun-exposed human skin and cutaneous carcinomas". Cancer. 94 (3): 814–9. doi:10.1002/cncr.10277. PMID 11857317.
  31. 1 2 Izquierdo JM, Majós N, Bonnal S, et al. (2005). "Regulation of Fas alternative splicing by antagonistic effects of TIA-1 and PTB on exon definition". Mol. Cell. 19 (4): 475–84. doi:10.1016/j.molcel.2005.06.015. PMID 16109372.
  32. 1 2 Zahler, A. M.; Damgaard, CK; Kjems, J; Caputi, M (2003). "SC35 and Heterogeneous Nuclear Ribonucleoprotein A/B Proteins Bind to a Juxtaposed Exonic Splicing Enhancer/Exonic Splicing Silencer Element to Regulate HIV-1 tat Exon 2 Splicing". Journal of Biological Chemistry. 279 (11): 10077–84. doi:10.1074/jbc.M312743200. PMID 14703516.
  33. Jacquenet, S.; Méreau, A; Bilodeau, PS; Damier, L; Stoltzfus, CM; Branlant, C (2001). "A Second Exon Splicing Silencer within Human Immunodeficiency Virus Type 1 tat Exon 2 Represses Splicing of Tat mRNA and Binds Protein hnRNP H". Journal of Biological Chemistry. 276 (44): 40464–75. doi:10.1074/jbc.M104070200. PMID 11526107.
  34. "HHMI Bulletin September 2005: Alternative Splicing". www.hhmi.org. Archived from the original on 22 June 2009. Retrieved 2009-05-26.
  35. Romero, Pedro; Zaidi, Saima; Fang, Ya Yin; Uversky, Vladimir; Dunker, Keith (2006). "Alternative splicing in concert with protein intrinsic disorder enables increased functional diversity in multicellular organisms.". Proc Natl Acad Sci U S A. 103: 8390–8395. Bibcode:2006PNAS..103.8390R. doi:10.1073/pnas.0507916103. PMC 1482503Freely accessible. PMID 16717195.
  36. Li, HD; Menon, R; Omenn, GS; Guan, Y (Jun 17, 2014). "The emerging era of genomic data integration for analyzing splice isoform function.". Trends in genetics : TIG. 30 (8): 340–347. doi:10.1016/j.tig.2014.05.005. PMID 24951248.
  37. 1 2 Eksi, R; Li, HD; Menon, R; Wen, Y; Omenn, GS; Kretzler, M; Guan, Y (Nov 2013). "Systematically differentiating functions for alternatively spliced isoforms through integrating RNA-seq data.". PLOS Computational Biology. 9 (11): e1003314. Bibcode:2013PLSCB...9E3314E. doi:10.1371/journal.pcbi.1003314. PMC 3820534Freely accessible. PMID 24244129.
  38. Irimia, Manuel; Rukov, Jakob; Penny, David; Roy, Scott (2007). "Functional and evolutionary analysis of alternatively spliced genes is consistent with an early eukaryotic origin of alternative splicing". BMC Evolutionary Biology. 7: 188. doi:10.1186/1471-2148-7-188. PMC 2082043Freely accessible. PMID 17916237.
  39. Ewing, B; Green P (June 2000). "Analysis of expressed sequence tags indicates 35,000 human genes". Nature Genetics. 25 (2): 232–234. doi:10.1038/76115. PMID 10835644.
  40. Crollius, HR; et al. (2000). "Estimate of human gene number provided by genome-wide analysis using Tetraodon nigroviridis DNA sequence". Nature Genetics. 25 (2): 235–238. doi:10.1038/76118. PMID 10835645.
  41. David Brett; Heike Pospisil; Juan Valcárcel; Jens Reich; Peer Bork (2002). "Alternative splicing and genome complexity". Nature Genetics. 30 (1): 29–30. doi:10.1038/ng803. PMID 11743582.
  42. Kim, E.; Magen, A.; Ast, G. (2006). "Different levels of alternative splicing among eukaryotes". Nucleic Acids Research. 35 (1): 125–31. doi:10.1093/nar/gkl924. PMC 1802581Freely accessible. PMID 17158149.
  43. López-Bigas, Núria; Audit, Benjamin; Ouzounis, Christos; Parra, Genís; Guigó, Roderic (2005). "Are splicing mutations the most frequent cause of hereditary disease?". FEBS Letters. 579 (9): 1900–3. doi:10.1016/j.febslet.2005.02.047. PMID 15792793.
  44. Ward, AJ; Cooper, TA (2010). "The pathobiology of splicing". J. Pathol. 220 (2): 152–163. doi:10.1002/path.2649. PMC 2855871Freely accessible. PMID 19918805.
  45. Omenn, GS; Guan, Y; Menon, R (May 3, 2014). "A New Class of Protein Cancer Biomarker Candidates: Differentially-Expressed Splice Variants of ERBB2 (HER2/neu) and ERBB1 (EGFR) in Breast Cancer Cell Lines.". Journal of proteomics. 107C: 103–112. doi:10.1016/j.jprot.2014.04.012. PMID 24802673.
  46. Sveen, A; Ågesen, TH; Nesbakken, A; Rognum, TO; Lothe, RA; Skotheim, RI (2011). "Transcriptome instability in colorectal cancer identified by exon microarray analyses: Associations with splicing factor expression levels and patient survival". Genome Medicine. 15: 672. doi:10.1186/1471-2164-15-672. PMC 3219073Freely accessible. PMID 25109687.
  47. Sveen, A; Johannessen, B; Teixeira, MR; Lothe, RA; Skotheim, RI (2014). "Transcriptome instability as a molecular pan-cancer characteristic of carcinomas". BMC Genomics. 3: 32. doi:10.1186/gm248. PMC 4137096Freely accessible. PMID 21619627.
  48. Kim E, Goren A, Ast G (2008). "Insights into the connection between cancer and alternative splicing". Trends Genet. 24 (1): 7–10. doi:10.1016/j.tig.2007.10.001. PMID 18054115.
  49. 1 2 Ghigna C, Giordano S, Shen H, et al. (2005). "Cell motility is controlled by SF2/ASF through alternative splicing of the Ron proto-oncogene". Mol. Cell. 20 (6): 881–90. doi:10.1016/j.molcel.2005.10.026. PMID 16364913.
  50. Hui L, Zhang X, Wu X, et al. (2004). "Identification of alternatively spliced mRNA variants related to cancers by genome-wide ESTs alignment". Oncogene. 23 (17): 3013–23. doi:10.1038/sj.onc.1207362. PMID 15048092.
  51. Danckwardt S, Neu-Yilik G, Thermann R, Frede U, Hentze MW, Kulozik AE (2002). "Abnormally spliced beta-globin mRNAs: a single point mutation generates transcripts sensitive and insensitive to nonsense-mediated mRNA decay". Blood. 99 (5): 1811–6. doi:10.1182/blood.V99.5.1811. PMID 11861299.
  52. Nestler EJ (December 2013). "Cellular basis of memory for addiction". Dialogues Clin. Neurosci. 15 (4): 431–443. PMC 3898681Freely accessible. PMID 24459410. DESPITE THE IMPORTANCE OF NUMEROUS PSYCHOSOCIAL FACTORS, AT ITS CORE, DRUG ADDICTION INVOLVES A BIOLOGICAL PROCESS: the ability of repeated exposure to a drug of abuse to induce changes in a vulnerable brain that drive the compulsive seeking and taking of drugs, and loss of control over drug use, that define a state of addiction. ... A large body of literature has demonstrated that such ΔFosB induction in D1-type NAc neurons increases an animal's sensitivity to drug as well as natural rewards and promotes drug self-administration, presumably through a process of positive reinforcement
  53. Ruffle JK (November 2014). "Molecular neurobiology of addiction: what's all the (Δ)FosB about?". Am J Drug Alcohol Abuse. 40 (6): 428–437. doi:10.3109/00952990.2014.933840. PMID 25083822. ΔFosB is an essential transcription factor implicated in the molecular and behavioral pathways of addiction following repeated drug exposure. The formation of ΔFosB in multiple brain regions, and the molecular pathway leading to the formation of AP-1 complexes is well understood. The establishment of a functional purpose for ΔFosB has allowed further determination as to some of the key aspects of its molecular cascades, involving effectors such as GluR2 (87,88), Cdk5 (93) and NFkB (100). Moreover, many of these molecular changes identified are now directly linked to the structural, physiological and behavioral changes observed following chronic drug exposure (60,95,97,102). New frontiers of research investigating the molecular roles of ΔFosB have been opened by epigenetic studies, and recent advances have illustrated the role of ΔFosB acting on DNA and histones, truly as a ‘‘molecular switch’’ (34). As a consequence of our improved understanding of ΔFosB in addiction, it is possible to evaluate the addictive potential of current medications (119), as well as use it as a biomarker for assessing the efficacy of therapeutic interventions (121,122,124). Some of these proposed interventions have limitations (125) or are in their infancy (75). However, it is hoped that some of these preliminary findings may lead to innovative treatments, which are much needed in addiction.
  54. Biliński P, Wojtyła A, Kapka-Skrzypczak L, Chwedorowicz R, Cyranka M, Studziński T (2012). "Epigenetic regulation in drug addiction". Ann. Agric. Environ. Med. 19 (3): 491–496. PMID 23020045. For these reasons, ΔFosB is considered a primary and causative transcription factor in creating new neural connections in the reward centre, prefrontal cortex, and other regions of the limbic system. This is reflected in the increased, stable and long-lasting level of sensitivity to cocaine and other drugs, and tendency to relapse even after long periods of abstinence. These newly constructed networks function very efficiently via new pathways as soon as drugs of abuse are further taken
  55. Olsen CM (December 2011). "Natural rewards, neuroplasticity, and non-drug addictions". Neuropharmacology. 61 (7): 1109–1122. doi:10.1016/j.neuropharm.2011.03.010. PMC 3139704Freely accessible. PMID 21459101.
  56. Luco, RF; Allo, M; Schor, IE; Kornblihtt, AR; Misteli, T. (2011). "Epigenetics in alternative pre-mRNA splicing". Cell. 144 (1): 16–26. doi:10.1016/j.cell.2010.11.056. PMC 3038581Freely accessible. PMID 21215366.
  57. Fairbrother, WG; Yeh, RF; Sharp, PA; Burge, CB (2002). "Predictive identification of exonic splicing enhancers in human genes". Science. 297 (5583): 1007–1013. Bibcode:2002Sci...297.1007F. doi:10.1126/science.1073774. PMID 12114529.
  58. Pan, Qun; Shai, Ofer; Misquitta, Christine; Zhang, Wen; Saltzman, Arneet L.; Mohammad, Naveed; Babak, Tomas; Siu, Henry; Hughes, Timothy R. (2004). "Revealing Global Regulatory Features of Mammalian Alternative Splicing Using a Quantitative Microarray Platform". Molecular Cell. 16 (6): 929–41. doi:10.1016/j.molcel.2004.12.004. PMID 15610736.
  59. Watkins, KH; Stewart, A; Fairbrother, WG (2009). "A rapid high-throughput method for mapping ribonucleoproteins (RNPs) on human pre-mRNA". J. Vis. Exp. 34: 1622. doi:10.3791/1622.
  60. Tuerk, C; Gold, L (1990). "Systematic evolution of ligands by exponential enrichment: RNA ligands to bacteriophage T4 DNA polymerase". Science. 249 (4968): 505–510. Bibcode:1990Sci...249..505T. doi:10.1126/science.2200121. PMID 2200121.
  61. Chang, B; Levin, J; Thompson, WA; Fairbrother, WG (2010). "High-throughput binding analysis determines the binding specificity of ASF/SF2 on alternatively spliced human pre-mRNAs". Comb. Chem. High Throughput Screen. 13 (3): 242–252. doi:10.2174/138620710790980522. PMID 20015017.
  62. Taggart, AJ; DeSimone, AM; Shih, JS; Filloux, ME; Fairbrother, WG (2012). "Large-scale mapping of branchpoints in human pre-mRNA transcripts in vivo". Nat. Struct. Mol. Biol. 19 (7): 719–721. doi:10.1038/nsmb.2327. PMC 3465671Freely accessible. PMID 22705790.

External links

Wikimedia Commons has media related to Alternative splicing.
This article is issued from Wikipedia - version of the 11/7/2016. The text is available under the Creative Commons Attribution/Share Alike but additional terms may apply for the media files.