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.
Previous studies have shown that protein-protein interactions among splicing factors may play an important role in pre-mRNA splicing. We report here identification and functional characterization of a new splicing factor, Sip1 (SC35-interacting protein 1). Sip1 was initially identified by virtue of its interaction with SC35, a splicing factor of the SR family. Sip1 interacts with not only several SR proteins but also with U1-70K and U2AF65, proteins associated with 5′ and 3′ splice sites, respectively. The predicted Sip1 sequence contains an arginine-serine-rich (RS) domain but does not have any known RNA-binding motifs, indicating that it is not a member of the SR family. Sip1 also contains a region with weak sequence similarity to the Drosophila splicing regulator suppressor of white apricot (SWAP). An essential role for Sip1 in pre-mRNA splicing was suggested by the observation that anti-Sip1 antibodies depleted splicing activity from HeLa nuclear extract. Purified recombinant Sip1 protein, but not other RS domain-containing proteins such as SC35, ASF/SF2, and U2AF65, restored the splicing activity of the Sip1-immunodepleted extract. Addition of U2AF65 protein further enhanced the splicing reconstitution by the Sip1 protein. Deficiency in the formation of both A and B splicing complexes in the Sip1-depleted nuclear extract indicates an important role of Sip1 in spliceosome assembly. Together, these results demonstrate that Sip1 is a novel RS domain-containing protein required for pre-mRNA splicing and that the functional role of Sip1 in splicing is distinct from those of known RS domain-containing splicing factors.Pre-mRNA splicing takes place in spliceosomes, the large RNA-protein complexes containing pre-mRNA, U1, U2, U4/6, and U5 small nuclear ribonucleoprotein particles (snRNPs), and a large number of accessory protein factors (for reviews, see references 21, 22, 37, 44, and 48). It is increasingly clear that the protein factors are important for pre-mRNA splicing and that studies of these factors are essential for further understanding of molecular mechanisms of pre-mRNA splicing.Most mammalian splicing factors have been identified by biochemical fractionation and purification (3, 15, 19, 3136, 45, 6971, 73), by using antibodies recognizing splicing factors (8, 9, 16, 17, 61, 66, 67, 74), and by sequence homology (25, 52, 74).Splicing factors containing arginine-serine-rich (RS) domains have emerged as important players in pre-mRNA splicing. These include members of the SR family, both subunits of U2 auxiliary factor (U2AF), and the U1 snRNP protein U1-70K (for reviews, see references 18, 41, and 59). Drosophila alternative splicing regulators transformer (Tra), transformer 2 (Tra2), and suppressor of white apricot (SWAP) also contain RS domains (20, 40, 42). RS domains in these proteins play important roles in pre-mRNA splicing (7, 71, 75), in nuclear localization of these splicing proteins (23, 40), and in protein-RNA interactions (56, 60, 64). Previous studies by us and others have demonstrated that one mechanism whereby SR proteins function in splicing is to mediate specific protein-protein interactions among spliceosomal components and between general splicing factors and alternative splicing regulators (1, 1a, 6, 10, 27, 63, 74, 77). Such protein-protein interactions may play critical roles in splice site recognition and association (for reviews, see references 4, 18, 37, 41, 47 and 59). Specific interactions among the splicing factors also suggest that it is possible to identify new splicing factors by their interactions with known splicing factors.Here we report identification of a new splicing factor, Sip1, by its interaction with the essential splicing factor SC35. The predicted Sip1 protein sequence contains an RS domain and a region with sequence similarity to the Drosophila splicing regulator, SWAP. We have expressed and purified recombinant Sip1 protein and raised polyclonal antibodies against the recombinant Sip1 protein. The anti-Sip1 antibodies specifically recognize a protein migrating at a molecular mass of approximately 210 kDa in HeLa nuclear extract. The anti-Sip1 antibodies sufficiently deplete Sip1 protein from the nuclear extract, and the Sip1-depleted extract is inactive in pre-mRNA splicing. Addition of recombinant Sip1 protein can partially restore splicing activity to the Sip1-depleted nuclear extract, indicating an essential role of Sip1 in pre-mRNA splicing. Other RS domain-containing proteins, including SC35, ASF/SF2, and U2AF65, cannot substitute for Sip1 in reconstituting splicing activity of the Sip1-depleted nuclear extract. However, addition of U2AF65 further increases splicing activity of Sip1-reconstituted nuclear extract, suggesting that there may be a functional interaction between Sip1 and U2AF65 in nuclear extract.  相似文献   

2.
3.
4.
Most human genes undergo alternative splicing, but aberrant splice forms are hallmarks of many cancers, usually resulting from mutations initiating abnormal exon skipping, intron retention, or the introduction of a new splice sites. We have identified a family of aberrant splice variants of HAS1 (the hyaluronan synthase 1 gene) in some B lineage cancers, characterized by exon skipping and/or partial intron retention events that occur either together or independently in different variants, apparently due to accumulation of inherited and acquired mutations. Cellular, biochemical, and oncogenic properties of full-length HAS1 (HAS1-FL) and HAS1 splice variants Va, Vb, and Vc (HAS1-Vs) are compared and characterized. When co-expressed, the properties of HAS1-Vs are dominant over those of HAS1-FL. HAS1-FL appears to be diffusely expressed in the cell, but HAS1-Vs are concentrated in the cytoplasm and/or Golgi apparatus. HAS1-Vs synthesize detectable de novo HA intracellularly. Each of the HAS1-Vs is able to relocalize HAS1-FL protein from diffuse cytoskeleton-anchored locations to deeper cytoplasmic spaces. This HAS1-Vs-mediated relocalization occurs through strong molecular interactions, which also serve to protect HAS1-FL from its otherwise high turnover kinetics. In co-transfected cells, HAS1-FL and HAS1-Vs interact with themselves and with each other to form heteromeric multiprotein assemblies. HAS1-Vc was found to be transforming in vitro and tumorigenic in vivo when introduced as a single oncogene to untransformed cells. The altered distribution and half-life of HAS1-FL, coupled with the characteristics of the HAS1-Vs suggest possible mechanisms whereby the aberrant splicing observed in human cancer may contribute to oncogenesis and disease progression.About 70–80% of human genes undergo alternative splicing, contributing to proteomic diversity and regulatory complexities in normal development (1). About 10% of mutations listed so far in the Human Gene Mutation Database (HGMD) of “gene lesions responsible for human inherited disease” were found to be located within splice sites. Furthermore, it is becoming increasingly apparent that aberrant splice variants, generated mostly due to splicing defects, play a key role in cancer. Germ line or acquired genomic changes (mutations) in/around splicing elements (24) promote aberrant splicing and aberrant protein isoforms.Hyaluronan (HA)3 is synthesized by three different plasma membrane-bound hyaluronan synthases (1, 2, and 3). HAS1 undergoes alternative and aberrant intronic splicing in multiple myeloma, producing truncated variants termed Va, Vb, and Vc (5, 6), which predicted for poor survival in a cohort of multiple myeloma patients (5). Our work suggests that this aberrant splicing arises due to inherited predispositions and acquired mutations in the HAS1 gene (7). Cancer-related, defective mRNA splicing caused by polymorphisms and/or mutations in splicing elements often results in inactivation of tumor suppressor activity (e.g. HRPT2 (8, 9), PTEN (10), MLHI (1114), and ATR (15)) or generation of dominant negative inhibitors (e.g. CHEK2 (16) and VWOX (17)). In breast cancer, aberrantly spliced forms of progesterone and estrogen receptors are found (reviewed in Ref. 3). Intronic mutations inactivate p53 through aberrant splicing and intron retention (18). Somatic mutations with the potential to alter splicing are frequent in some cancers (1925). Single nucleotide polymorphisms in the cyclin D1 proto-oncogene predispose to aberrant splicing and the cyclin D1b intronic splice variant (2629). Cyclin D1b confers anchorage independence, is tumorogenic in vivo, and is detectable in human tumors (30), but as yet no clinical studies have confirmed an impact on outcome. On the other hand, aberrant splicing of HAS1 shows an association between aberrant splice variants and malignancy, suggesting that such variants may be potential therapeutic targets and diagnostic indicators (19, 3133). Increased HA expression has been associated with malignant progression of multiple tumor types, including breast, prostate, colon, glioma, mesothelioma, and multiple myeloma (34). The three mammalian HA synthase (HAS) isoenzymes synthesize HA and are integral transmembrane proteins with a probable porelike structural assembly (3539). Although in humans, the three HAS genes are located on different chromosomes (hCh19, hCh8, and hCh16, respectively) (40), they share a high degree of sequence homology (41, 42). HAS isoenzymes synthesize a different size range of HA molecules, which exhibit different functions (43, 44). HASs contribute to a variety of cancers (4555). Overexpression of HASs promotes growth and/or metastatic development in fibrosarcoma, prostate, and mammary carcinoma, and the removal of the HA matrix from a migratory cell membrane inhibits cell movement (45, 53). HAS2 confers anchorage independence (56). Our work has shown aberrant HAS1 splicing in multiple myeloma (5) and Waldenstrom''s macroglobulinemia (6). HAS1 is overexpressed in colon (57), ovarian (58), endometrial (59), mesothelioma (60), and bladder cancers (61). A HAS1 splice variant is detected in bladder cancer (61).Here, we characterize molecular and biochemical characteristics of HAS1 variants (HAS1-Vs) (5), generated by aberrant splicing. Using transient transfectants and tagged HAS1 family constructs, we show that HAS1-Vs differ in cellular localization, de novo HA localization, and turnover kinetics, as compared with HAS1-FL, and dominantly influence HAS1-FL when co-expressed. HAS1-Vs proteins form intra- and intermolecular associations among themselves and with HAS1-FL, including covalent interactions and multimer formation. HAS1-Vc supports vigorous cellular transformation of NIH3T3 cells in vitro, and HAS1-Vc-transformed NIH3T3 cells are tumorogenic in vivo.  相似文献   

5.
6.
7.
Alternative splicing is regulated by cis sequences in the pre-mRNA that serve as binding sites for trans-acting alternative splicing factors. In a previous study, we used bioinformatics and molecular biology to identify and confirm that the intronic hexamer sequence UCUAUC is a nematode alternative splicing regulatory element. In this study, we used RNA affinity chromatography to identify trans factors that bind to this sequence. HRP-2, the Caenorhabditis elegans homolog of human heterogeneous nuclear ribonucleoproteins Q and R, binds to UCUAUC in the context of unc-52 intronic regulatory sequences as well as to RNAs containing tandem repeats of this sequence. The three Us in the hexamer are the most important determinants of this binding specificity. We demonstrate, using RNA interference, that HRP-2 regulates the alternative splicing of two genes, unc-52 and lin-10, both of which have cassette exons flanked by an intronic UCUAUC motif. We propose that HRP-2 is a protein responsible for regulating alternative splicing through binding interactions with the UCUAUC sequence.Alternative pre-mRNA splicing is a mechanism for generating multiple mRNA isoforms from a single gene. This process can allow a gene to encode for more than one protein isoform. For some genes, it is a mechanism for regulating message stability through production of alternative mRNA isoforms that are substrates for the nonsense-mediated mRNA decay pathway (1). The majority of human genes undergo alternative splicing (2), and the process can be regulated in tissue-specific and developmental stage-specific manners. Current models propose that cis elements on the pre-mRNA, in exons and introns, serve as recognition sites for trans-acting protein factors that bind to the pre-mRNA and regulate assembly of the splicing machinery, thus regulating splice site choice (3).In recent years, a number of groups have employed bioinformatics techniques to identify cis splicing regulatory elements (4). These techniques include using multiple interspecies sequence alignments to identify conserved intronic regions, identification of short sequences in exons that are bounded by weak consensus splice sites, and identification of common intronic sequences flanking similarly regulated alternative exons (59). These efforts have added many new sequences to the list of known and potential splicing regulators. The identification of the protein factor partners for these sequences will be important for understanding their function in alternative splicing regulation.Experimental approaches have identified alternative splicing factors that interact with specific cis elements (10), but the number of trans factors discovered still lags behind the number of newly identified cis element partners. Some examples of well-characterized cis element/trans-acting factor interactions include the NOVA K homology domain splicing factor binding to the sequence UCAY (11), the FOX splicing factors binding to the sequence UGCAUG (1214), and hnRNP3 F/H proteins binding to the sequence GGGG (15, 16). By using cross-linking immunoprecipitation followed by large scale sequencing, entire catalogs of RNAs that the splicing factors NOVA, SF2/ASF, and FOX2 bind to in vivo have been determined (1719). These approaches have led to models for how the proteins binding to their cis regulatory elements may alter splicing. These models include a role for the relative position of a cis element to an alternative cassette exon in determining alternative exon inclusion or skipping (18, 19).In a previous bioinformatics analysis of evolutionarily conserved intronic sequences flanking alternatively spliced exons, we identified the hexamer sequence UCUAUC as a novel splicing regulatory element (8). UCUAUC is found flanking both sides of alternative exon 16 of the unc-52 gene of Caenorhabditis elegans. Genetic analysis of a class of viable unc-52 mutants led to the discovery that exons 16–18 are alternative cassette exons and that every combination of skipping and inclusion of these three exons occurs (20). This splicing is regulated by the alternative splicing factor MEC-8 (21). Fig. 1A shows a schematic diagram of the alternatively spliced region of unc-52, with the MEC-8-enhanced alternative splicing events indicated. Using an unc-52 splicing reporter trans gene containing alternative exons 15–19, we previously reported that alternative splicing is regulated by the intronic motif UCUAUC in the intron downstream of exon 16 (8). In addition we showed that this element works cooperatively with a UGCAUG hexamer (the consensus FOX-1-binding site) in the upstream intron to regulate alternative splicing (8).Open in a separate windowFIGURE 1.RNA affinity chromatography identifies HRP-2 as binding to UCUAUC elements. A, schematic representation of the alternatively spliced region of unc-52 (adapted from Ref. 21). The alternative splicing events promoted by MEC-8 are indicated by bold lines. The lines next to introns 15 and 16 are the sites of the UCUAUC elements in those introns whose sequences were used in the RNA affinity chromatography. B, table showing sequences of RNAs immobilized to beads in the RNA affinity chromatography experiment. C, Coomassie-stained SDS-PAGE analysis of RNA affinity chromatography. C. elegans embryo extract was incubated with the different immobilized RNA substrates listed on top of the gel. Proteins identified by mass spectrometry are listed to the right of the gel, with arrows pointing to coincident protein bands. D, the left panel shows the silver stain result for the RNA affinity chromatography experiment. Each lane represents a different immobilized substrate, as indicated above. The band corresponding to HRP-2 is indicated by an arrow. The right panel is an immunoblot of the same gel using anti-HRP-2 polyclonal antibody. E, anti-HRP-2 immunoblot of an RNA affinity chromatography experiment for the indicated substrates.In this study, we report the results of a biochemical identification of a protein factor from C. elegans that binds to the UCUAUC intronic splicing regulatory element. We transcribed different short RNA sequences containing the UCUAUC element in its native intronic context, or as part of a repeating unit, and immobilized these onto agarose beads. After passing embryo extracts across these beads, we found that the protein HRP-2, the C. elegans homolog of the mammalian hnRNP Q/R proteins, binds to this sequence with high affinity. By using RNAi to reduce the level of HRP-2 in worms, we observed changes in alternative splicing of unc-52 and lin-10, two genes that contain UCUAUC elements in introns flanking alternative exons. We propose that HRP-2 is an alternative splicing factor that works through the UCUAUC intronic elements to regulate alternative splicing.  相似文献   

8.
The Bcl-x pre-mRNA is alternatively spliced to produce the anti-apoptotic Bcl-xL and the pro-apoptotic Bcl-xS isoforms. By performing deletion mutagenesis on a human Bcl-x minigene, we have identified a novel exonic element that controls the use of the 5′ splice site of Bcl-xS. The proximal portion of this element acts as a repressor and is located downstream of an enhancer. Further mutational analysis provided a detailed topological map of the regulatory activities revealing a sharp transition between enhancer and repressor sequences. Portions of the enhancer can function when transplanted in another alternative splicing unit. Chromatography and immunoprecipitation assays indicate that the silencer element interacts with heterogeneous ribonucleoprotein particle (hnRNP) K, consistent with the presence of putative high affinity sites for this protein. Finally, down-regulation of hnRNP K by RNA interference enhanced splicing to Bcl-xS, an effect seen only when the sequences bound by hnRNP K are present. Our results therefore document a clear role for hnRNP K in preventing the production of the pro-apoptotic Bcl-xS splice isoform.Alternative splicing is a major mechanism used to augment the number of proteins encoded by the genome. It is estimated that as many as 97% of multiple exon pre-mRNAs undergo alternative splicing (1, 2). Disruption of alternative splicing by mutating important regulatory sequences or by altering the expression or activity of proteins controlling splice site selection has been linked with different diseases, including cancer (37). Apoptosis is an important and complex cellular program involved in development and differentiation in higher organisms (8, 9). However, its aberrant control often contributes to cancer development and the resistance of cancer cells to drug therapy (1013).Genes implicated in the apoptotic pathway are alternatively spliced often to produce protein isoforms with distinct or even antagonistic activities (14, 15). A good example is the apoptotic regulator Bcl-x, which is alternatively spliced to produce two major isoforms, the anti-apoptotic Bcl-xL protein and the shorter pro-apoptotic Bcl-xS isoform (16). This alternative splicing decision involves a competition between two 5′ splice sites; the use of the downstream site creates Bcl-xL, and the use of the upstream one produces Bcl-xS (Fig. 1A). Bcl-xL is always the predominant form in cancer cells, and overexpressing it can confer resistance to chemotherapeutic agents (1722). On the other hand, overexpression of the pro-apoptotic Bcl-xS isoform enhances sensitivity to the topoisomerase inhibitor etoposide and to taxol in a breast cancer cell line, while triggering apoptosis in melanoma cell lines (23, 24). Using antisense technologies to improve the production of the Bcl-xS splice variant can also induce apoptosis in cancer cells (2527).Open in a separate windowFIGURE 1.A, alternative splicing of Bcl-x produces two major isoforms, Bcl-xL and Bcl-xS. B, regulation of Bcl-x alternative splicing. The enhancer elements are shown as white boxes, and the repressors are black. The pointed and flat arrows indicate positive and negative regulation, respectively. Protein kinase C inhibition relieves repression caused by the SB1 element on the Bcl-xS splice site (36). The repressor elements CRCE1, recognized by SAP155, and CRCE2 mediate the production of Bcl-xS by ceramide as when induced by gemcitabine in A549 cells (38, 39). hnRNP F/H binds to the B2G element to enhance the production of the Bcl-xS isoform (41). RBM25, through an element located upstream of the Bcl-xS splice site, can also augment its use (44). A large intronic region (IRE) mediates the Bcl-xL increase caused by interleukin-6 (IL-6), granulocyte-macrophage colony-stimulating factor (GM-CSF), and 12-O-tetradecanoylphorbol-13-acetate (TPA) (35). Finally, the B3 region also enhances Bcl-xL formation through the binding of SRp30c to AM2 and ML2 and the U1 snRNP to two cryptic 5′ splice sites (42).Alternative splicing is regulated by different proteins bound to sequence elements near splice sites. A variety of mechanisms is used to achieve regulation. Some splicing factors act by recruiting or inhibiting the binding of different components of the spliceosome. Others may change the conformation of the pre-mRNA to mask a splice site or to bring a pair of splice sites into closer proximity (28, 29).Although individual factors can have a strong and specific effect on splicing decisions, alternative splicing often relies on a combination of factors to determine the appropriate levels of isoforms. The implication of multiple proteins likely provides additional levels of regulation that helps attuned splicing control to a variety of stresses, environmental cues, and growth conditions. In several cases, the interaction of regulatory factors can be antagonistic. For example, in the Drosophila male-specific-lethal-2 (msl-2) pre-mRNA, recruitment of SXL to a uridine-rich region interferes with the binding of TIA-1 that is necessary for efficient U1 snRNP2 recruitment at the 5′ splice site (30). On the same pre-mRNA, SXL also diminishes U2AF recognition of the polypyrimidine tract at the 3′ splice site. TIA proteins bound to a U-rich element on the avian myosin phosphatase targeting subunit-1 (MYPT1) pre-mRNA repress the binding of PTB (31). PTB can also reduce the recruitment of ETR-3 to intronic elements near exon 5 of cardiac troponin T (32). In neurons, the binding of PTB to the introns surrounding the N1 exon of c-src is antagonized by nPTB protein, promoting exon inclusion. On the hnRNP A1 pre-mRNA, PTB diminishes the binding of SRp30c to the intronic CE9 element, reducing the inhibition by this protein on the use of the downstream 3′ splice site (33). SC35 and hnRNP A1 have partially overlapping binding sites on the human immunodeficiency virus 1 (HIV-1) tat exon 2. Preferential binding of SC35 enhances the inclusion of the exon, whereas hnRNP A1, by reducing SC35 binding, increases exclusion (34). Thus, the competition provided by an overlapping or a closely abutting pair of enhancer/ silencer represents a simple and frequent mechanism of splicing control.The regulation of Bcl-x alternative splicing has received some attention in recent years leading to the discovery of several cis-acting elements and a few trans-acting control factors (Fig. 1B). Intronic regions downstream from the Bcl-xL 5′ splice site have been implicated as mediating signals from cytokines such as interleukin-6 and granulocyte-macrophage colony-stimulating factor (35). In addition, we have reported that an element located 187 nt upstream of the Bcl-xS splice site mediates a protein kinase C-dependent signal that represses splicing to the Bcl-xS donor site (36). On the other hand, ceramide enhances the use of the Bcl-xS 5′ splice site by lifting the repression mediated by two other elements (37, 38). The activity of one of these apparently involves SAP155 (39). The RNA-binding protein Sam68, under the control of the tyrosine kinase Fyn, can also increase the production of Bcl-xS in cooperation with hnRNP A1 (40), and this effect is inhibited by overexpression of ASF/SF2. The Bcl-x sequences bound by the above factors remain to be identified. We also uncovered enhancer elements for Bcl-xS and Bcl-xL. hnRNP F and H bind downstream of the Bcl-xS 5′ splice site to stimulate splicing to that site (41). Enhancement of Bcl-xL is conferred by SRp30c, which binds upstream of the 5′ splice site to antagonize the repressor activity of pseudo 5′ splice sites (42). Recently, the SR protein SC35 was shown to increase the production of Bcl-xS (43). Finally, the binding of RBM25 to a sequence element upstream of the Bcl-xS 5′ splice site stimulated its use, possibly by recruiting U1 snRNP through its interaction with the U1-associated protein hLuc7A (44). Thus, the region located between the two competing 5′ splice sites of Bcl-x is densely populated by splicing control elements.In this study, we have pursued our characterization of Bcl-x splicing control by examining the contribution of sequences directly upstream of the Bcl-xS donor site. Our mutational approach identified a region containing flanking enhancer and silencer activities. The activity of the repressor portion is mediated by hnRNP K, which makes this protein an anti-apoptotic regulator.  相似文献   

9.
10.
11.
12.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

13.
14.
15.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

16.
17.
18.
19.
Mycobacterium leprae, which has undergone reductive evolution leaving behind a minimal set of essential genes, has retained intervening sequences in four of its genes implicating a vital role for them in the survival of the leprosy bacillus. A single in-frame intervening sequence has been found embedded within its recA gene. Comparison of the M. leprae recA intervening sequence with the known intervening sequences indicated that it has the consensus amino acid sequence necessary for being a LAGLIDADG-type homing endonuclease. In light of massive gene decay and function loss in the leprosy bacillus, we sought to investigate whether its recA intervening sequence encodes a catalytically active homing endonuclease. Here we show that the purified M. leprae RecA intein (PI-MleI) binds to cognate DNA and displays endonuclease activity in the presence of alternative divalent cations, Mg2+ or Mn2+. A combination of approaches, including four complementary footprinting assays such as DNase I, copper-phenanthroline, methylation protection, and KMnO4, enhancement of 2-aminopurine fluorescence, and mapping of the cleavage site revealed that PI-MleI binds to cognate DNA flanking its insertion site, induces helical distortion at the cleavage site, and generates two staggered double strand breaks. Taken together, these results implicate that PI-MleI possesses a modular structure with separate domains for DNA target recognition and cleavage, each with distinct sequence preferences. From a biological standpoint, it is tempting to speculate that our findings have implications for understanding the evolution of the LAGLIDADG family of homing endonucleases.Mycobacterium leprae, a Gram-positive rod-shaped bacillus, mostly found in warm tropical countries, is the bacterium that causes leprosy in humans (1). The lack of understanding of the basic biology of M. leprae is believed to be the key factor for the failure of leprosy research to advance. The genome sequence of M. leprae contains 3.27 Mb and has an average G + C content of 57.8%, values much lower than the corresponding values for Mycobacterium tuberculosis, which are ∼4.41 Mb and 65.6% G + C, respectively (2). There are some 1500 genes that are common to both M. leprae and M. tuberculosis. The comparative genome analysis suggests that both species of mycobacteria are derived from a common ancestor and, at one stage, had gene pools of similar size. The downsizing of the M. tuberculosis genome from ∼4.41 to 3.27 Mb of M. leprae would account for the loss of some 1200 protein-coding sequences (1, 3). There is evidence that many of the genes that were present in the genome of M. leprae have truly been lost (1, 3). Comparative genomics of M. leprae with that of M. tuberculosis indicate that the former has undergone substantial downsizing, losing more than 2000 genes, thus suggesting an extreme case of reductive evolution in a microbial pathogen (1). With the availability of the M. leprae genome sequence, using functional genomics approaches, it is possible to identify the gene products, elucidate the mechanism of their action, and identify novel drug targets for rational design of new therapeutic regimens and drugs to treat leprosy.Eubacterial RecA proteins catalyze a set of biochemical reactions that are essential for homologous recombination, DNA repair, restoration of stalled replication forks, and SOS response (47). RecA protein and the process of homologous recombination, which is the main mechanism of genetic exchange, are evolutionarily conserved among a range of organisms (4, 7). Perhaps the most striking development in the field of RecA protein biology was the discovery of an in-frame insertion of an intein-coding sequence in the recA genes of M. tuberculosis and M. leprae (8, 9). In these organisms, RecA is synthesized as a large precursor, which undergoes protein splicing to excise the intein, and the two flanking domains called exteins are ligated together to generate a functionally active RecA protein (9, 10). The milieu in which RecA precursor undergoes splicing differs substantially between M. tuberculosis and M. leprae. M. leprae RecA precursor (79 kDa) undergoes splicing only in mycobacterial species, whereas M. tuberculosis RecA precursor (85 kDa) is spliced efficiently in Escherichia coli as well (911). Intriguingly, M. tuberculosis and M. leprae RecA inteins differ greatly in their size, primary sequence, and location within the recA gene, thereby suggesting two independent origins during evolution (9). The occurrence of inteins in the obligate mycobacterial pathogens, M. tuberculosis, M. leprae, and Mycobacterium microti, suggested that RecA inteins might play a role in mycobacterial functions related to pathogenesis or virulence (9). Previously, we have shown that M. tuberculosis RecA intein (PI-MtuI),2 which contains Walker A motif, displays dual target specificity in the presence of alternative cofactors in an ATP-dependent manner (12, 13).Since their discovery in Saccharomyces cerevisiae (14, 15), a large number of putative homing endonucleases have been found in a diverse range of proteins in all the three domains of life (1619). The majority of inteins possess the protein splicing and homing endonuclease activities (18, 19). Homing endonucleases are a class of diverse rare-cutting enzymes that promote site-specific transposition of their encoding genetic elements by inflicting double-stranded DNA breaks via different cleavage mechanisms in alleles lacking these elements (1823). In addition, these are characterized by their ability to bind long DNA target sites (14–40 bp), and their tolerance of minor sequence changes in their binding region. These have been divided into highly divergent subfamilies on the basis of conserved sequence and structural motifs as follows: LAGLIDADG, GIY-YIG, HNH, His-Cys box, and the more recently identified PD(D/E)XK families (1824). LAGLIDADG homing enzymes, which include the largest family, contain one or two copies of the conserved dodecapeptide motif and utilize an extended protein-DNA interface covering up to 40 bp to acquire their necessary specificity (1822). The LAGLIDADG sequence is a part of the conserved 10- or 12-residue sequence motif defining the family of LAGLIDADG-type homing endonucleases; therefore, it is designated as deca- or dodecapeptide motif (19).Comparison of the M. leprae recA intervening sequence with known intervening sequences indicated that it has the consensus amino acid sequence necessary for being a LAGLIDADG-type homing endonuclease (25, 26). In light of massive gene decay and function loss in the leprosy bacillus, and dissimilarities in size and primary structures among mycobacterial inteins, we sought to investigate whether M. leprae recA intervening sequence encodes a catalytically active homing endonuclease. In this study, we show that the purified M. leprae RecA intein (PI-MleI) binds to cognate DNA and displays endonuclease activity in the presence of alternative divalent cations Mg2+ or Mn2+. Furthermore, using a variety of approaches, we have mapped the positions of PI-MleI binding as well as cleavage in the cognate DNA, thus providing the most comprehensive analysis of PI-MleI. Taken together, these results suggest that PI-MleI possesses a modular structure with functionally separable domains for DNA target recognition and cleavage, each with distinct sequence preferences. These results provide insights into understanding the function and evolution of the family of LAGLIDADG homing endonucleases.  相似文献   

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