@article {49595, title = {Developmental expression of chicken FOXN1 and putative target genes during feather development.}, volume = {58}, year = {2014}, month = {2014}, pages = {57-64}, abstract = {

FOXN1 is a member of the forkhead box family of transcription factors. FOXN1 is crucial for hair outgrowth and thymus differentiation in mammals. Unlike the thymus, which is found in all amniotes, hair is an epidermal appendage that arose after the last shared common ancestor between mammals and birds, and hair and feathers differ markedly in their differentiation and gene expression. Here, we show that FOXN1 is expressed in embryonic chicken feathers, nails and thymus, demonstrating an evolutionary conservation that goes beyond obvious homology. At embryonic day (ED) 12, FOXN1 is expressed in some feather buds and at ED13 expression extends along the length of the feather filament. At ED14 FOXN1 mRNA is restricted to the proximal feather filament and is not detectable in distal feather shafts. At the base of the feather, FOXN1 is expressed in the epithelium of the feather sheath and distal barb and marginal plate, whereas in the midsection FOXN1 transcripts are mainly detected in the barb plates of the feather filament. FOXN1 is also expressed in claws; however, no expression was detected in skin or scales. Despite expression of FOXN1 in developing feathers, examination of chick homologs of five putative mammalian FOXN1 target genes shows that, while these genes are expressed in feathers, there is little similarity to the FOXN1 expression pattern, suggesting that some gene regulatory networks may have diverged during evolution of epidermal appendages.

}, keywords = {Amino Acid Sequence, Animals, Biological Evolution, Blotting, Western, Cell Differentiation, Cells, Cultured, Chick Embryo, Chickens, Cloning, Molecular, Embryo, Nonmammalian, Epidermis, Feathers, Forkhead Transcription Factors, Gene Expression Regulation, Developmental, In Situ Hybridization, Molecular Sequence Data, Morphogenesis, Phylogeny, Real-Time Polymerase Chain Reaction, Reverse Transcriptase Polymerase Chain Reaction, RNA, Messenger, Sequence Homology, Amino Acid}, issn = {1696-3547}, doi = {10.1387/ijdb.130023sy}, author = {Darnell, Diana K and Zhang, Li S and Hannenhalli, Sridhar and Yaklichkin, Sergey Y} } @article {49736, title = {Stoichiometry of site-specific lysine acetylation in an entire proteome.}, journal = {J Biol Chem}, volume = {289}, year = {2014}, month = {2014 Aug 1}, pages = {21326-38}, abstract = {

Acetylation of lysine ϵ-amino groups influences many cellular processes and has been mapped to thousands of sites across many organisms. Stoichiometric information of acetylation is essential to accurately interpret biological significance. Here, we developed and employed a novel method for directly quantifying stoichiometry of site-specific acetylation in the entire proteome of Escherichia coli. By coupling isotopic labeling and a novel pairing algorithm, our approach performs an in silico enrichment of acetyl peptides, circumventing the need for immunoenrichment. We investigated the function of the sole NAD(+)-dependent protein deacetylase, CobB, on both site-specific and global acetylation. We quantified 2206 peptides from 899 proteins and observed a wide distribution of acetyl stoichiometry, ranging from less than 1\% up to 98\%. Bioinformatic analysis revealed that metabolic enzymes, which either utilize or generate acetyl-CoA, and proteins involved in transcriptional and translational processes displayed the highest degree of acetylation. Loss of CobB led to increased global acetylation at low stoichiometry sites and induced site-specific changes at high stoichiometry sites, and biochemical analysis revealed altered acetyl-CoA metabolism. Thus, this study demonstrates that sirtuin deacetylase deficiency leads to both site-specific and global changes in protein acetylation stoichiometry, affecting central metabolism.

}, keywords = {Acetylation, Amino Acid Sequence, Bacterial Proteins, Chromatography, High Pressure Liquid, Computational Biology, Escherichia coli, Lysine, Molecular Sequence Data, Proteome, Tandem Mass Spectrometry}, issn = {1083-351X}, doi = {10.1074/jbc.M114.581843}, author = {Baeza, Josue and Dowell, James A and Smallegan, Michael J and Fan, Jing and Amador-Noguez, Daniel and Khan, Zia and Denu, John M} } @article {49775, title = {Archaeosortases and exosortases are widely distributed systems linking membrane transit with posttranslational modification.}, journal = {J Bacteriol}, volume = {194}, year = {2012}, month = {2012 Jan}, pages = {36-48}, abstract = {

Multiple new prokaryotic C-terminal protein-sorting signals were found that reprise the tripartite architecture shared by LPXTG and PEP-CTERM: motif, TM helix, basic cluster. Defining hidden Markov models were constructed for all. PGF-CTERM occurs in 29 archaeal species, some of which have more than 50 proteins that share the domain. PGF-CTERM proteins include the major cell surface protein in Halobacterium, a glycoprotein with a partially characterized diphytanylglyceryl phosphate linkage near its C terminus. Comparative genomics identifies a distant exosortase homolog, designated archaeosortase A (ArtA), as the likely protein-processing enzyme for PGF-CTERM. Proteomics suggests that the PGF-CTERM region is removed. Additional systems include VPXXXP-CTERM/archeaosortase B in two of the same archaea and PEF-CTERM/archaeosortase C in four others. Bacterial exosortases often fall into subfamilies that partner with very different cohorts of extracellular polymeric substance biosynthesis proteins; several species have multiple systems. Variant systems include the VPDSG-CTERM/exosortase C system unique to certain members of the phylum Verrucomicrobia, VPLPA-CTERM/exosortase D in several alpha- and deltaproteobacterial species, and a dedicated (single-target) VPEID-CTERM/exosortase E system in alphaproteobacteria. Exosortase-related families XrtF in the class Flavobacteria and XrtG in Gram-positive bacteria mark distinctive conserved gene neighborhoods. A picture emerges of an ancient and now well-differentiated superfamily of deeply membrane-embedded protein-processing enzymes. Their target proteins are destined to transit cellular membranes during their biosynthesis, during which most undergo additional posttranslational modifications such as glycosylation.

}, keywords = {Amino Acid Sequence, Aminoacyltransferases, Archaeal Proteins, Bacterial Proteins, Cell Membrane, Cysteine Endopeptidases, Gene Expression Regulation, Archaeal, Gene Expression Regulation, Bacterial, Gene Expression Regulation, Enzymologic, Molecular Sequence Data, Protein Processing, Post-Translational}, issn = {1098-5530}, doi = {10.1128/JB.06026-11}, author = {Haft, Daniel H and Payne, Samuel H and Selengut, Jeremy D} } @article {38119, title = {Archaeosortases and exosortases are widely distributed systems linking membrane transit with posttranslational modification}, journal = {Journal of bacteriologyJournal of bacteriology}, volume = {194}, year = {2012}, note = {http://www.ncbi.nlm.nih.gov/pubmed/22037399?dopt=Abstract}, type = {10.1128/JB.06026-11}, abstract = {Multiple new prokaryotic C-terminal protein-sorting signals were found that reprise the tripartite architecture shared by LPXTG and PEP-CTERM: motif, TM helix, basic cluster. Defining hidden Markov models were constructed for all. PGF-CTERM occurs in 29 archaeal species, some of which have more than 50 proteins that share the domain. PGF-CTERM proteins include the major cell surface protein in Halobacterium, a glycoprotein with a partially characterized diphytanylglyceryl phosphate linkage near its C terminus. Comparative genomics identifies a distant exosortase homolog, designated archaeosortase A (ArtA), as the likely protein-processing enzyme for PGF-CTERM. Proteomics suggests that the PGF-CTERM region is removed. Additional systems include VPXXXP-CTERM/archeaosortase B in two of the same archaea and PEF-CTERM/archaeosortase C in four others. Bacterial exosortases often fall into subfamilies that partner with very different cohorts of extracellular polymeric substance biosynthesis proteins; several species have multiple systems. Variant systems include the VPDSG-CTERM/exosortase C system unique to certain members of the phylum Verrucomicrobia, VPLPA-CTERM/exosortase D in several alpha- and deltaproteobacterial species, and a dedicated (single-target) VPEID-CTERM/exosortase E system in alphaproteobacteria. Exosortase-related families XrtF in the class Flavobacteria and XrtG in Gram-positive bacteria mark distinctive conserved gene neighborhoods. A picture emerges of an ancient and now well-differentiated superfamily of deeply membrane-embedded protein-processing enzymes. Their target proteins are destined to transit cellular membranes during their biosynthesis, during which most undergo additional posttranslational modifications such as glycosylation.}, keywords = {Amino Acid Sequence, Aminoacyltransferases, Archaeal Proteins, Bacterial Proteins, Cell Membrane, Cysteine Endopeptidases, Gene Expression Regulation, Archaeal, Gene Expression Regulation, Bacterial, Gene Expression Regulation, Enzymologic, Molecular Sequence Data, Protein Processing, Post-Translational}, author = {Haft, Daniel H. and Payne, Samuel H. and J. Selengut} } @article {49531, title = {Plasmodium falciparum merozoite surface protein 1 blocks the proinflammatory protein S100P.}, volume = {109}, year = {2012}, month = {2012 Apr 3}, pages = {5429-34}, abstract = {

The malaria parasite, Plasmodium falciparum, and the human immune system have coevolved to ensure that the parasite is not eliminated and reinfection is not resisted. This relationship is likely mediated through a myriad of host-parasite interactions, although surprisingly few such interactions have been identified. Here we show that the 33-kDa fragment of P. falciparum merozoite surface protein 1 (MSP1(33)), an abundant protein that is shed during red blood cell invasion, binds to the proinflammatory protein, S100P. MSP1(33) blocks S100P-induced NFκB activation in monocytes and chemotaxis in neutrophils. Remarkably, S100P binds to both dimorphic alleles of MSP1, estimated to have diverged >27 Mya, suggesting an ancient, conserved relationship between these parasite and host proteins that may serve to attenuate potentially damaging inflammatory responses.

}, keywords = {Amino Acid Sequence, Animals, Calcium-Binding Proteins, Chromatography, Gel, Electrophoresis, Polyacrylamide Gel, Enzyme-Linked Immunosorbent Assay, HUMANS, Merozoite Surface Protein 1, Microscopy, Confocal, Molecular Sequence Data, Neoplasm Proteins, Plasmodium falciparum, Sequence Homology, Amino Acid, Surface Plasmon Resonance}, issn = {1091-6490}, doi = {10.1073/pnas.1202689109}, author = {Waisberg, Michael and Cerqueira, Gustavo C and Yager, Stephanie B and Francischetti, Ivo M B and Lu, Jinghua and Gera, Nidhi and Srinivasan, Prakash and Miura, Kazutoyo and Rada, Balazs and Lukszo, Jan and Barbian, Kent D and Leto, Thomas L and Porcella, Stephen F and Narum, David L and El-Sayed, Najib and Miller, Louis H and Pierce, Susan K} } @article {38573, title = {Whole genome analysis of Leptospira licerasiae provides insight into leptospiral evolution and pathogenicity}, journal = {PLoS neglected tropical diseasesPLoS neglected tropical diseases}, volume = {6}, year = {2012}, note = {http://www.ncbi.nlm.nih.gov/pubmed/23145189?dopt=Abstract}, type = {10.1371/journal.pntd.0001853}, abstract = {The whole genome analysis of two strains of the first intermediately pathogenic leptospiral species to be sequenced (Leptospira licerasiae strains VAR010 and MMD0835) provides insight into their pathogenic potential and deepens our understanding of leptospiral evolution. Comparative analysis of eight leptospiral genomes shows the existence of a core leptospiral genome comprising 1547 genes and 452 conserved genes restricted to infectious species (including L. licerasiae) that are likely to be pathogenicity-related. Comparisons of the functional content of the genomes suggests that L. licerasiae retains several proteins related to nitrogen, amino acid and carbohydrate metabolism which might help to explain why these Leptospira grow well in artificial media compared with pathogenic species. L. licerasiae strains VAR010(T) and MMD0835 possess two prophage elements. While one element is circular and shares homology with LE1 of L. biflexa, the second is cryptic and homologous to a previously identified but unnamed region in L. interrogans serovars Copenhageni and Lai. We also report a unique O-antigen locus in L. licerasiae comprised of a 6-gene cluster that is unexpectedly short compared with L. interrogans in which analogous regions may include >90 such genes. Sequence homology searches suggest that these genes were acquired by lateral gene transfer (LGT). Furthermore, seven putative genomic islands ranging in size from 5 to 36 kb are present also suggestive of antecedent LGT. How Leptospira become naturally competent remains to be determined, but considering the phylogenetic origins of the genes comprising the O-antigen cluster and other putative laterally transferred genes, L. licerasiae must be able to exchange genetic material with non-invasive environmental bacteria. The data presented here demonstrate that L. licerasiae is genetically more closely related to pathogenic than to saprophytic Leptospira and provide insight into the genomic bases for its infectiousness and its unique antigenic characteristics.}, keywords = {DNA, Bacterial, Evolution, Molecular, Gene Transfer, Horizontal, Genome, Bacterial, Genomic islands, HUMANS, Leptospira, Molecular Sequence Data, Multigene Family, Prophages, Sequence Analysis, DNA, Virulence factors}, author = {Ricaldi, Jessica N. and Fouts, Derrick E. and J. Selengut and Harkins, Derek M. and Patra, Kailash P. and Moreno, Angelo and Lehmann, Jason S. and Purushe, Janaki and Sanka, Ravi and Torres, Michael and Webster, Nicholas J. and Vinetz, Joseph M. and Matthias, Michael A.} } @article {49776, title = {Whole genome analysis of Leptospira licerasiae provides insight into leptospiral evolution and pathogenicity.}, journal = {PLoS Negl Trop Dis}, volume = {6}, year = {2012}, month = {2012}, pages = {e1853}, abstract = {

The whole genome analysis of two strains of the first intermediately pathogenic leptospiral species to be sequenced (Leptospira licerasiae strains VAR010 and MMD0835) provides insight into their pathogenic potential and deepens our understanding of leptospiral evolution. Comparative analysis of eight leptospiral genomes shows the existence of a core leptospiral genome comprising 1547 genes and 452 conserved genes restricted to infectious species (including L. licerasiae) that are likely to be pathogenicity-related. Comparisons of the functional content of the genomes suggests that L. licerasiae retains several proteins related to nitrogen, amino acid and carbohydrate metabolism which might help to explain why these Leptospira grow well in artificial media compared with pathogenic species. L. licerasiae strains VAR010(T) and MMD0835 possess two prophage elements. While one element is circular and shares homology with LE1 of L. biflexa, the second is cryptic and homologous to a previously identified but unnamed region in L. interrogans serovars Copenhageni and Lai. We also report a unique O-antigen locus in L. licerasiae comprised of a 6-gene cluster that is unexpectedly short compared with L. interrogans in which analogous regions may include >90 such genes. Sequence homology searches suggest that these genes were acquired by lateral gene transfer (LGT). Furthermore, seven putative genomic islands ranging in size from 5 to 36 kb are present also suggestive of antecedent LGT. How Leptospira become naturally competent remains to be determined, but considering the phylogenetic origins of the genes comprising the O-antigen cluster and other putative laterally transferred genes, L. licerasiae must be able to exchange genetic material with non-invasive environmental bacteria. The data presented here demonstrate that L. licerasiae is genetically more closely related to pathogenic than to saprophytic Leptospira and provide insight into the genomic bases for its infectiousness and its unique antigenic characteristics.

}, keywords = {DNA, Bacterial, Evolution, Molecular, Gene Transfer, Horizontal, Genome, Bacterial, Genomic islands, HUMANS, Leptospira, Molecular Sequence Data, Multigene Family, Prophages, Sequence Analysis, DNA, Virulence factors}, issn = {1935-2735}, doi = {10.1371/journal.pntd.0001853}, author = {Ricaldi, Jessica N and Fouts, Derrick E and Selengut, Jeremy D and Harkins, Derek M and Patra, Kailash P and Moreno, Angelo and Lehmann, Jason S and Purushe, Janaki and Sanka, Ravi and Torres, Michael and Webster, Nicholas J and Vinetz, Joseph M and Matthias, Michael A} } @article {49744, title = {Accurate proteome-wide protein quantification from high-resolution 15N mass spectra.}, journal = {Genome Biol}, volume = {12}, year = {2011}, month = {2011}, pages = {R122}, abstract = {

In quantitative mass spectrometry-based proteomics, the metabolic incorporation of a single source of 15N-labeled nitrogen has many advantages over using stable isotope-labeled amino acids. However, the lack of a robust computational framework for analyzing the resulting spectra has impeded wide use of this approach. We have addressed this challenge by introducing a new computational methodology for analyzing 15N spectra in which quantification is integrated with identification. Application of this method to an Escherichia coli growth transition reveals significant improvement in quantification accuracy over previous methods.

}, keywords = {algorithms, Amino Acid Sequence, Bacterial Proteins, Escherichia coli, Isotope Labeling, Mass Spectrometry, Molecular Sequence Data, Nitrogen Isotopes, Proteome, proteomics, Sensitivity and Specificity, software}, issn = {1474-760X}, doi = {10.1186/gb-2011-12-12-r122}, author = {Khan, Zia and Amini, Sasan and Bloom, Joshua S and Ruse, Cristian and Caudy, Amy A and Kruglyak, Leonid and Singh, Mona and Perlman, David H and Tavazoie, Saeed} } @article {49779, title = {Sites Inferred by Metabolic Background Assertion Labeling (SIMBAL): adapting the Partial Phylogenetic Profiling algorithm to scan sequences for signatures that predict protein function.}, journal = {BMC Bioinformatics}, volume = {11}, year = {2010}, month = {2010}, pages = {52}, abstract = {

BACKGROUND: Comparative genomics methods such as phylogenetic profiling can mine powerful inferences from inherently noisy biological data sets. We introduce Sites Inferred by Metabolic Background Assertion Labeling (SIMBAL), a method that applies the Partial Phylogenetic Profiling (PPP) approach locally within a protein sequence to discover short sequence signatures associated with functional sites. The approach is based on the basic scoring mechanism employed by PPP, namely the use of binomial distribution statistics to optimize sequence similarity cutoffs during searches of partitioned training sets.

RESULTS: Here we illustrate and validate the ability of the SIMBAL method to find functionally relevant short sequence signatures by application to two well-characterized protein families. In the first example, we partitioned a family of ABC permeases using a metabolic background property (urea utilization). Thus, the TRUE set for this family comprised members whose genome of origin encoded a urea utilization system. By moving a sliding window across the sequence of a permease, and searching each subsequence in turn against the full set of partitioned proteins, the method found which local sequence signatures best correlated with the urea utilization trait. Mapping of SIMBAL "hot spots" onto crystal structures of homologous permeases reveals that the significant sites are gating determinants on the cytosolic face rather than, say, docking sites for the substrate-binding protein on the extracellular face. In the second example, we partitioned a protein methyltransferase family using gene proximity as a criterion. In this case, the TRUE set comprised those methyltransferases encoded near the gene for the substrate RF-1. SIMBAL identifies sequence regions that map onto the substrate-binding interface while ignoring regions involved in the methyltransferase reaction mechanism in general. Neither method for training set construction requires any prior experimental characterization.

CONCLUSIONS: SIMBAL shows that, in functionally divergent protein families, selected short sequences often significantly outperform their full-length parent sequence for making functional predictions by sequence similarity, suggesting avenues for improved functional classifiers. When combined with structural data, SIMBAL affords the ability to localize and model functional sites.

}, keywords = {algorithms, Amino Acid Sequence, Gene Expression Profiling, Molecular Sequence Data, Phylogeny, Proteins, Sequence Analysis, Protein, Structure-Activity Relationship}, issn = {1471-2105}, doi = {10.1186/1471-2105-11-52}, author = {Selengut, Jeremy D and Rusch, Douglas B and Haft, Daniel H} } @article {38506, title = {Sites Inferred by Metabolic Background Assertion Labeling (SIMBAL): adapting the Partial Phylogenetic Profiling algorithm to scan sequences for signatures that predict protein function}, journal = {BMC bioinformaticsBMC Bioinformatics}, volume = {11}, year = {2010}, note = {http://www.ncbi.nlm.nih.gov/pubmed/20102603?dopt=Abstract}, type = {10.1186/1471-2105-11-52}, abstract = {BACKGROUND: Comparative genomics methods such as phylogenetic profiling can mine powerful inferences from inherently noisy biological data sets. We introduce Sites Inferred by Metabolic Background Assertion Labeling (SIMBAL), a method that applies the Partial Phylogenetic Profiling (PPP) approach locally within a protein sequence to discover short sequence signatures associated with functional sites. The approach is based on the basic scoring mechanism employed by PPP, namely the use of binomial distribution statistics to optimize sequence similarity cutoffs during searches of partitioned training sets. RESULTS: Here we illustrate and validate the ability of the SIMBAL method to find functionally relevant short sequence signatures by application to two well-characterized protein families. In the first example, we partitioned a family of ABC permeases using a metabolic background property (urea utilization). Thus, the TRUE set for this family comprised members whose genome of origin encoded a urea utilization system. By moving a sliding window across the sequence of a permease, and searching each subsequence in turn against the full set of partitioned proteins, the method found which local sequence signatures best correlated with the urea utilization trait. Mapping of SIMBAL "hot spots" onto crystal structures of homologous permeases reveals that the significant sites are gating determinants on the cytosolic face rather than, say, docking sites for the substrate-binding protein on the extracellular face. In the second example, we partitioned a protein methyltransferase family using gene proximity as a criterion. In this case, the TRUE set comprised those methyltransferases encoded near the gene for the substrate RF-1. SIMBAL identifies sequence regions that map onto the substrate-binding interface while ignoring regions involved in the methyltransferase reaction mechanism in general. Neither method for training set construction requires any prior experimental characterization. CONCLUSIONS: SIMBAL shows that, in functionally divergent protein families, selected short sequences often significantly outperform their full-length parent sequence for making functional predictions by sequence similarity, suggesting avenues for improved functional classifiers. When combined with structural data, SIMBAL affords the ability to localize and model functional sites.}, keywords = {algorithms, Amino Acid Sequence, Gene Expression Profiling, Molecular Sequence Data, Phylogeny, Proteins, Sequence Analysis, Protein, Structure-Activity Relationship}, author = {J. Selengut and Rusch, Douglas B. and Haft, Daniel H.} } @article {49778, title = {Unexpected abundance of coenzyme F(420)-dependent enzymes in Mycobacterium tuberculosis and other actinobacteria.}, journal = {J Bacteriol}, volume = {192}, year = {2010}, month = {2010 Nov}, pages = {5788-98}, abstract = {

Regimens targeting Mycobacterium tuberculosis, the causative agent of tuberculosis (TB), require long courses of treatment and a combination of three or more drugs. An increase in drug-resistant strains of M. tuberculosis demonstrates the need for additional TB-specific drugs. A notable feature of M. tuberculosis is coenzyme F(420), which is distributed sporadically and sparsely among prokaryotes. This distribution allows for comparative genomics-based investigations. Phylogenetic profiling (comparison of differential gene content) based on F(420) biosynthesis nominated many actinobacterial proteins as candidate F(420)-dependent enzymes. Three such families dominated the results: the luciferase-like monooxygenase (LLM), pyridoxamine 5{\textquoteright}-phosphate oxidase (PPOX), and deazaflavin-dependent nitroreductase (DDN) families. The DDN family was determined to be limited to F(420)-producing species. The LLM and PPOX families were observed in F(420)-producing species as well as species lacking F(420) but were particularly numerous in many actinobacterial species, including M. tuberculosis. Partitioning the LLM and PPOX families based on an organism{\textquoteright}s ability to make F(420) allowed the application of the SIMBAL (sites inferred by metabolic background assertion labeling) profiling method to identify F(420)-correlated subsequences. These regions were found to correspond to flavonoid cofactor binding sites. Significantly, these results showed that M. tuberculosis carries at least 28 separate F(420)-dependent enzymes, most of unknown function, and a paucity of flavin mononucleotide (FMN)-dependent proteins in these families. While prevalent in mycobacteria, markers of F(420) biosynthesis appeared to be absent from the normal human gut flora. These findings suggest that M. tuberculosis relies heavily on coenzyme F(420) for its redox reactions. This dependence and the cofactor{\textquoteright}s rarity may make F(420)-related proteins promising drug targets.

}, keywords = {Actinobacteria, Amino Acid Sequence, Binding Sites, Coenzymes, Flavonoids, Gene Expression Profiling, Gene Expression Regulation, Bacterial, Genome, Bacterial, molecular biology, Molecular Sequence Data, Molecular Structure, Mycobacterium tuberculosis, Phylogeny, Protein Conformation, Riboflavin}, issn = {1098-5530}, doi = {10.1128/JB.00425-10}, author = {Selengut, Jeremy D and Haft, Daniel H} } @article {38556, title = {Unexpected abundance of coenzyme F(420)-dependent enzymes in Mycobacterium tuberculosis and other actinobacteria}, journal = {Journal of bacteriologyJournal of bacteriology}, volume = {192}, year = {2010}, note = {http://www.ncbi.nlm.nih.gov/pubmed/20675471?dopt=Abstract}, type = {10.1128/JB.00425-10}, abstract = {Regimens targeting Mycobacterium tuberculosis, the causative agent of tuberculosis (TB), require long courses of treatment and a combination of three or more drugs. An increase in drug-resistant strains of M. tuberculosis demonstrates the need for additional TB-specific drugs. A notable feature of M. tuberculosis is coenzyme F(420), which is distributed sporadically and sparsely among prokaryotes. This distribution allows for comparative genomics-based investigations. Phylogenetic profiling (comparison of differential gene content) based on F(420) biosynthesis nominated many actinobacterial proteins as candidate F(420)-dependent enzymes. Three such families dominated the results: the luciferase-like monooxygenase (LLM), pyridoxamine 5{\textquoteright}-phosphate oxidase (PPOX), and deazaflavin-dependent nitroreductase (DDN) families. The DDN family was determined to be limited to F(420)-producing species. The LLM and PPOX families were observed in F(420)-producing species as well as species lacking F(420) but were particularly numerous in many actinobacterial species, including M. tuberculosis. Partitioning the LLM and PPOX families based on an organism{\textquoteright}s ability to make F(420) allowed the application of the SIMBAL (sites inferred by metabolic background assertion labeling) profiling method to identify F(420)-correlated subsequences. These regions were found to correspond to flavonoid cofactor binding sites. Significantly, these results showed that M. tuberculosis carries at least 28 separate F(420)-dependent enzymes, most of unknown function, and a paucity of flavin mononucleotide (FMN)-dependent proteins in these families. While prevalent in mycobacteria, markers of F(420) biosynthesis appeared to be absent from the normal human gut flora. These findings suggest that M. tuberculosis relies heavily on coenzyme F(420) for its redox reactions. This dependence and the cofactor{\textquoteright}s rarity may make F(420)-related proteins promising drug targets.}, keywords = {Actinobacteria, Amino Acid Sequence, Binding Sites, Coenzymes, Flavonoids, Gene Expression Profiling, Gene Expression Regulation, Bacterial, Genome, Bacterial, molecular biology, Molecular Sequence Data, Molecular Structure, Mycobacterium tuberculosis, Phylogeny, Protein Conformation, Riboflavin}, author = {J. Selengut and Haft, Daniel H.} } @article {49646, title = {The genome of the blood fluke Schistosoma mansoni.}, journal = {Nature}, volume = {460}, year = {2009}, month = {2009 Jul 16}, pages = {352-8}, abstract = {

Schistosoma mansoni is responsible for the neglected tropical disease schistosomiasis that affects 210 million people in 76 countries. Here we present analysis of the 363 megabase nuclear genome of the blood fluke. It encodes at least 11,809 genes, with an unusual intron size distribution, and new families of micro-exon genes that undergo frequent alternative splicing. As the first sequenced flatworm, and a representative of the Lophotrochozoa, it offers insights into early events in the evolution of the animals, including the development of a body pattern with bilateral symmetry, and the development of tissues into organs. Our analysis has been informed by the need to find new drug targets. The deficits in lipid metabolism that make schistosomes dependent on the host are revealed, and the identification of membrane receptors, ion channels and more than 300 proteases provide new insights into the biology of the life cycle and new targets. Bioinformatics approaches have identified metabolic chokepoints, and a chemogenomic screen has pinpointed schistosome proteins for which existing drugs may be active. The information generated provides an invaluable resource for the research community to develop much needed new control tools for the treatment and eradication of this important and neglected disease.

}, keywords = {Animals, Biological Evolution, Exons, Genes, Helminth, Genome, Helminth, Host-Parasite Interactions, Introns, Molecular Sequence Data, Physical Chromosome Mapping, Schistosoma mansoni, Schistosomiasis mansoni}, issn = {1476-4687}, doi = {10.1038/nature08160}, author = {Berriman, Matthew and Haas, Brian J and LoVerde, Philip T and Wilson, R Alan and Dillon, Gary P and Cerqueira, Gustavo C and Mashiyama, Susan T and Al-Lazikani, Bissan and Andrade, Luiza F and Ashton, Peter D and Aslett, Martin A and Bartholomeu, Daniella C and Blandin, Ga{\"e}lle and Caffrey, Conor R and Coghlan, Avril and Coulson, Richard and Day, Tim A and Delcher, Art and DeMarco, Ricardo and Djikeng, Appolinaire and Eyre, Tina and Gamble, John A and Ghedin, Elodie and Gu, Yong and Hertz-Fowler, Christiane and Hirai, Hirohisha and Hirai, Yuriko and Houston, Robin and Ivens, Alasdair and Johnston, David A and Lacerda, Daniela and Macedo, Camila D and McVeigh, Paul and Ning, Zemin and Oliveira, Guilherme and Overington, John P and Parkhill, Julian and Pertea, Mihaela and Pierce, Raymond J and Protasio, Anna V and Quail, Michael A and Rajandream, Marie-Ad{\`e}le and Rogers, Jane and Sajid, Mohammed and Salzberg, Steven L and Stanke, Mario and Tivey, Adrian R and White, Owen and Williams, David L and Wortman, Jennifer and Wu, Wenjie and Zamanian, Mostafa and Zerlotini, Adhemar and Fraser-Liggett, Claire M and Barrell, Barclay G and El-Sayed, Najib M} } @article {49644, title = {Genomic organization and expression profile of the mucin-associated surface protein (masp) family of the human pathogen Trypanosoma cruzi.}, journal = {Nucleic Acids Res}, volume = {37}, year = {2009}, month = {2009 Jun}, pages = {3407-17}, abstract = {

A novel large multigene family was recently identified in the human pathogen Trypanosoma cruzi, causative agent of Chagas disease, and corresponds to approximately 6\% of the parasite diploid genome. The predicted gene products, mucin-associated surface proteins (MASPs), are characterized by highly conserved N- and C-terminal domains and a strikingly variable and repetitive central region. We report here an analysis of the genomic organization and expression profile of masp genes. Masps are not randomly distributed throughout the genome but instead are clustered with genes encoding mucin and other surface protein families. Masp transcripts vary in size, are preferentially expressed during the trypomastigote stage and contain highly conserved 5{\textquoteright} and 3{\textquoteright} untranslated regions. A sequence analysis of a trypomastigote cDNA library reveals the expression of multiple masp variants with a bias towards a particular masp subgroup. Immunofluorescence assays using antibodies generated against a MASP peptide reveals that the expression of particular MASPs at the cell membrane is limited to subsets of the parasite population. Western blots of phosphatidylinositol-specific phospholipase C (PI-PLC)-treated parasites suggest that MASP may be GPI-anchored and shed into the medium culture, thus contributing to the large repertoire of parasite polypeptides that are exposed to the host immune system.

}, keywords = {3{\textquoteright} Flanking Region, 5{\textquoteright} Flanking Region, Amino Acid Sequence, Animals, Base Sequence, Conserved Sequence, Gene Expression Profiling, Genes, Protozoan, Genome, Protozoan, Membrane Proteins, Molecular Sequence Data, Mucins, Multigene Family, Protozoan Proteins, RNA, Messenger, Trypanosoma cruzi}, issn = {1362-4962}, doi = {10.1093/nar/gkp172}, author = {Bartholomeu, Daniella C and Cerqueira, Gustavo C and Le{\~a}o, Ana Carolina A and daRocha, Wanderson D and Pais, Fabiano S and Macedo, Camila and Djikeng, Appolinaire and Teixeira, Santuza M R and El-Sayed, Najib M} } @article {38528, title = {Three genomes from the phylum Acidobacteria provide insight into the lifestyles of these microorganisms in soils}, journal = {Applied and environmental microbiologyApplied and environmental microbiology}, volume = {75}, year = {2009}, note = {http://www.ncbi.nlm.nih.gov/pubmed/19201974?dopt=Abstract}, type = {10.1128/AEM.02294-08}, abstract = {The complete genomes of three strains from the phylum Acidobacteria were compared. Phylogenetic analysis placed them as a unique phylum. They share genomic traits with members of the Proteobacteria, the Cyanobacteria, and the Fungi. The three strains appear to be versatile heterotrophs. Genomic and culture traits indicate the use of carbon sources that span simple sugars to more complex substrates such as hemicellulose, cellulose, and chitin. The genomes encode low-specificity major facilitator superfamily transporters and high-affinity ABC transporters for sugars, suggesting that they are best suited to low-nutrient conditions. They appear capable of nitrate and nitrite reduction but not N(2) fixation or denitrification. The genomes contained numerous genes that encode siderophore receptors, but no evidence of siderophore production was found, suggesting that they may obtain iron via interaction with other microorganisms. The presence of cellulose synthesis genes and a large class of novel high-molecular-weight excreted proteins suggests potential traits for desiccation resistance, biofilm formation, and/or contribution to soil structure. Polyketide synthase and macrolide glycosylation genes suggest the production of novel antimicrobial compounds. Genes that encode a variety of novel proteins were also identified. The abundance of acidobacteria in soils worldwide and the breadth of potential carbon use by the sequenced strains suggest significant and previously unrecognized contributions to the terrestrial carbon cycle. Combining our genomic evidence with available culture traits, we postulate that cells of these isolates are long-lived, divide slowly, exhibit slow metabolic rates under low-nutrient conditions, and are well equipped to tolerate fluctuations in soil hydration.}, keywords = {Anti-Bacterial Agents, bacteria, Biological Transport, Carbohydrate Metabolism, Cyanobacteria, DNA, Bacterial, Fungi, Genome, Bacterial, Macrolides, Molecular Sequence Data, Nitrogen, Phylogeny, Proteobacteria, Sequence Analysis, DNA, Sequence Homology, Soil Microbiology}, author = {Ward, Naomi L. and Challacombe, Jean F. and Janssen, Peter H. and Henrissat, Bernard and Coutinho, Pedro M. and Wu, Martin and Xie, Gary and Haft, Daniel H. and Sait, Michelle and Badger, Jonathan and Barabote, Ravi D. and Bradley, Brent and Brettin, Thomas S. and Brinkac, Lauren M. and Bruce, David and Creasy, Todd and Daugherty, Sean C. and Davidsen, Tanja M. and DeBoy, Robert T. and Detter, J. Chris and Dodson, Robert J. and Durkin, A. Scott and Ganapathy, Anuradha and Gwinn-Giglio, Michelle and Han, Cliff S. and Khouri, Hoda and Kiss, Hajnalka and Kothari, Sagar P. and Madupu, Ramana and Nelson, Karen E. and Nelson, William C. and Paulsen, Ian and Penn, Kevin and Ren, Qinghu and Rosovitz, M. J. and J. Selengut and Shrivastava, Susmita and Sullivan, Steven A. and Tapia, Roxanne and Thompson, L. Sue and Watkins, Kisha L. and Yang, Qi and Yu, Chunhui and Zafar, Nikhat and Zhou, Liwei and Kuske, Cheryl R.} } @article {49780, title = {Three genomes from the phylum Acidobacteria provide insight into the lifestyles of these microorganisms in soils.}, journal = {Appl Environ Microbiol}, volume = {75}, year = {2009}, month = {2009 Apr}, pages = {2046-56}, abstract = {

The complete genomes of three strains from the phylum Acidobacteria were compared. Phylogenetic analysis placed them as a unique phylum. They share genomic traits with members of the Proteobacteria, the Cyanobacteria, and the Fungi. The three strains appear to be versatile heterotrophs. Genomic and culture traits indicate the use of carbon sources that span simple sugars to more complex substrates such as hemicellulose, cellulose, and chitin. The genomes encode low-specificity major facilitator superfamily transporters and high-affinity ABC transporters for sugars, suggesting that they are best suited to low-nutrient conditions. They appear capable of nitrate and nitrite reduction but not N(2) fixation or denitrification. The genomes contained numerous genes that encode siderophore receptors, but no evidence of siderophore production was found, suggesting that they may obtain iron via interaction with other microorganisms. The presence of cellulose synthesis genes and a large class of novel high-molecular-weight excreted proteins suggests potential traits for desiccation resistance, biofilm formation, and/or contribution to soil structure. Polyketide synthase and macrolide glycosylation genes suggest the production of novel antimicrobial compounds. Genes that encode a variety of novel proteins were also identified. The abundance of acidobacteria in soils worldwide and the breadth of potential carbon use by the sequenced strains suggest significant and previously unrecognized contributions to the terrestrial carbon cycle. Combining our genomic evidence with available culture traits, we postulate that cells of these isolates are long-lived, divide slowly, exhibit slow metabolic rates under low-nutrient conditions, and are well equipped to tolerate fluctuations in soil hydration.

}, keywords = {Anti-Bacterial Agents, bacteria, Biological Transport, Carbohydrate Metabolism, Cyanobacteria, DNA, Bacterial, Fungi, Genome, Bacterial, Macrolides, Molecular Sequence Data, Nitrogen, Phylogeny, Proteobacteria, Sequence Analysis, DNA, Sequence Homology, Soil Microbiology}, issn = {1098-5336}, doi = {10.1128/AEM.02294-08}, author = {Ward, Naomi L and Challacombe, Jean F and Janssen, Peter H and Henrissat, Bernard and Coutinho, Pedro M and Wu, Martin and Xie, Gary and Haft, Daniel H and Sait, Michelle and Badger, Jonathan and Barabote, Ravi D and Bradley, Brent and Brettin, Thomas S and Brinkac, Lauren M and Bruce, David and Creasy, Todd and Daugherty, Sean C and Davidsen, Tanja M and DeBoy, Robert T and Detter, J Chris and Dodson, Robert J and Durkin, A Scott and Ganapathy, Anuradha and Gwinn-Giglio, Michelle and Han, Cliff S and Khouri, Hoda and Kiss, Hajnalka and Kothari, Sagar P and Madupu, Ramana and Nelson, Karen E and Nelson, William C and Paulsen, Ian and Penn, Kevin and Ren, Qinghu and Rosovitz, M J and Selengut, Jeremy D and Shrivastava, Susmita and Sullivan, Steven A and Tapia, Roxanne and Thompson, L Sue and Watkins, Kisha L and Yang, Qi and Yu, Chunhui and Zafar, Nikhat and Zhou, Liwei and Kuske, Cheryl R} } @article {49675, title = {Two alternatively spliced isoforms of the Arabidopsis SR45 protein have distinct roles during normal plant development.}, journal = {Plant Physiol}, volume = {150}, year = {2009}, month = {2009 Jul}, pages = {1450-8}, abstract = {

The serine-arginine-rich (SR) proteins constitute a conserved family of pre-mRNA splicing factors. In Arabidopsis (Arabidopsis thaliana), they are encoded by 19 genes, most of which are themselves alternatively spliced. In the case of SR45, the use of alternative 3{\textquoteright} splice sites 21 nucleotides apart generates two alternatively spliced isoforms. Isoform 1 (SR45.1) has an insertion relative to isoform 2 (SR45.2) that replaces a single arginine with eight amino acids (TSPQRKTG). The biological implications of SR45 alternative splicing have been unclear. A previously described loss-of-function mutant affecting both isoforms, sr45-1, shows several developmental defects, including defects in petal development and root growth. We found that the SR45 promoter is highly active in regions with actively growing and dividing cells. We also tested the ability of each SR45 isoform to complement the sr45-1 mutant by overexpression of isoform-specific green fluorescent protein (GFP) fusion proteins. As expected, transgenic plants overexpressing either isoform displayed both nuclear speckles and GFP fluorescence throughout the nucleoplasm. We found that SR45.1-GFP complements the flower petal phenotype, but not the root growth phenotype. Conversely, SR45.2-GFP complements root growth but not floral morphology. Mutation of a predicted phosphorylation site within the alternatively spliced segment, SR45.1-S219A-GFP, does not affect complementation. However, a double mutation affecting both serine-219 and the adjacent threonine-218 (SR45.1-T218A + S219A-GFP) behaves like isoform 2, complementing the root but not the floral phenotype. In conclusion, our study provides evidence that the two alternatively spliced isoforms of SR45 have distinct biological functions.

}, keywords = {Alternative Splicing, Amino Acid Sequence, Arabidopsis, Arabidopsis Proteins, Carrier Proteins, Flowers, Molecular Sequence Data, Mutation, Plant Roots, Protein Isoforms, Ribonucleoproteins, RNA-Binding Proteins, sequence alignment}, issn = {0032-0889}, doi = {10.1104/pp.109.138180}, author = {Zhang, Xiao-Ning and Mount, Stephen M} } @article {49676, title = {The draft genome of the transgenic tropical fruit tree papaya (Carica papaya Linnaeus).}, journal = {Nature}, volume = {452}, year = {2008}, month = {2008 Apr 24}, pages = {991-6}, abstract = {

Papaya, a fruit crop cultivated in tropical and subtropical regions, is known for its nutritional benefits and medicinal applications. Here we report a 3x draft genome sequence of {\textquoteright}SunUp{\textquoteright} papaya, the first commercial virus-resistant transgenic fruit tree to be sequenced. The papaya genome is three times the size of the Arabidopsis genome, but contains fewer genes, including significantly fewer disease-resistance gene analogues. Comparison of the five sequenced genomes suggests a minimal angiosperm gene set of 13,311. A lack of recent genome duplication, atypical of other angiosperm genomes sequenced so far, may account for the smaller papaya gene number in most functional groups. Nonetheless, striking amplifications in gene number within particular functional groups suggest roles in the evolution of tree-like habit, deposition and remobilization of starch reserves, attraction of seed dispersal agents, and adaptation to tropical daylengths. Transgenesis at three locations is closely associated with chloroplast insertions into the nuclear genome, and with topoisomerase I recognition sites. Papaya offers numerous advantages as a system for fruit-tree functional genomics, and this draft genome sequence provides the foundation for revealing the basis of Carica{\textquoteright}s distinguishing morpho-physiological, medicinal and nutritional properties.

}, keywords = {Arabidopsis, Carica, Contig Mapping, Databases, Genetic, Genes, Plant, Genome, Plant, Molecular Sequence Data, Plants, Genetically Modified, sequence alignment, Sequence Analysis, DNA, Transcription Factors, Tropical Climate}, issn = {1476-4687}, doi = {10.1038/nature06856}, author = {Ming, Ray and Hou, Shaobin and Feng, Yun and Yu, Qingyi and Dionne-Laporte, Alexandre and Saw, Jimmy H and Senin, Pavel and Wang, Wei and Ly, Benjamin V and Lewis, Kanako L T and Salzberg, Steven L and Feng, Lu and Jones, Meghan R and Skelton, Rachel L and Murray, Jan E and Chen, Cuixia and Qian, Wubin and Shen, Junguo and Du, Peng and Eustice, Moriah and Tong, Eric and Tang, Haibao and Lyons, Eric and Paull, Robert E and Michael, Todd P and Wall, Kerr and Rice, Danny W and Albert, Henrik and Wang, Ming-Li and Zhu, Yun J and Schatz, Michael and Nagarajan, Niranjan and Acob, Ricelle A and Guan, Peizhu and Blas, Andrea and Wai, Ching Man and Ackerman, Christine M and Ren, Yan and Liu, Chao and Wang, Jianmei and Wang, Jianping and Na, Jong-Kuk and Shakirov, Eugene V and Haas, Brian and Thimmapuram, Jyothi and Nelson, David and Wang, Xiyin and Bowers, John E and Gschwend, Andrea R and Delcher, Arthur L and Singh, Ratnesh and Suzuki, Jon Y and Tripathi, Savarni and Neupane, Kabi and Wei, Hairong and Irikura, Beth and Paidi, Maya and Jiang, Ning and Zhang, Wenli and Presting, Gernot and Windsor, Aaron and Navajas-P{\'e}rez, Rafael and Torres, Manuel J and Feltus, F Alex and Porter, Brad and Li, Yingjun and Burroughs, A Max and Luo, Ming-Cheng and Liu, Lei and Christopher, David A and Mount, Stephen M and Moore, Paul H and Sugimura, Tak and Jiang, Jiming and Schuler, Mary A and Friedman, Vikki and Mitchell-Olds, Thomas and Shippen, Dorothy E and dePamphilis, Claude W and Palmer, Jeffrey D and Freeling, Michael and Paterson, Andrew H and Gonsalves, Dennis and Wang, Lei and Alam, Maqsudul} } @article {49642, title = {Members of a large retroposon family are determinants of post-transcriptional gene expression in Leishmania.}, journal = {PLoS Pathog}, volume = {3}, year = {2007}, month = {2007 Sep 7}, pages = {1291-307}, abstract = {

Trypanosomatids are unicellular protists that include the human pathogens Leishmania spp. (leishmaniasis), Trypanosoma brucei (sleeping sickness), and Trypanosoma cruzi (Chagas disease). Analysis of their recently completed genomes confirmed the presence of non-long-terminal repeat retrotransposons, also called retroposons. Using the 79-bp signature sequence common to all trypanosomatid retroposons as bait, we identified in the Leishmania major genome two new large families of small elements--LmSIDER1 (785 copies) and LmSIDER2 (1,073 copies)--that fulfill all the characteristics of extinct trypanosomatid retroposons. LmSIDERs are approximately 70 times more abundant in L. major compared to T. brucei and are found almost exclusively within the 3{\textquoteright}-untranslated regions (3{\textquoteright}UTRs) of L. major mRNAs. We provide experimental evidence that LmSIDER2 act as mRNA instability elements and that LmSIDER2-containing mRNAs are generally expressed at lower levels compared to the non-LmSIDER2 mRNAs. The considerable expansion of LmSIDERs within 3{\textquoteright}UTRs in an organism lacking transcriptional control and their role in regulating mRNA stability indicate that Leishmania have probably recycled these short retroposons to globally modulate the expression of a number of genes. To our knowledge, this is the first example in eukaryotes of the domestication and expansion of a family of mobile elements that have evolved to fulfill a critical cellular function.

}, keywords = {3{\textquoteright} Untranslated Regions, Animals, Base Sequence, Biological Evolution, Down-Regulation, Gene Expression Regulation, Genome, Protozoan, Leishmania, Leishmania major, Molecular Sequence Data, Retroelements, RNA, Messenger, sequence alignment, Trypanosoma brucei brucei, Trypanosoma cruzi}, issn = {1553-7374}, doi = {10.1371/journal.ppat.0030136}, author = {Bringaud, Frederic and M{\"u}ller, Michaela and Cerqueira, Gustavo Coutinho and Smith, Martin and Rochette, Annie and el-Sayed, Najib M A and Papadopoulou, Barbara and Ghedin, Elodie} } @article {49681, title = {Spliceosomal small nuclear RNA genes in 11 insect genomes.}, journal = {RNA}, volume = {13}, year = {2007}, month = {2007 Jan}, pages = {5-14}, abstract = {

The removal of introns from the primary transcripts of protein-coding genes is accomplished by the spliceosome, a large macromolecular complex of which small nuclear RNAs (snRNAs) are crucial components. Following the recent sequencing of the honeybee (Apis mellifera) genome, we used various computational methods, ranging from sequence similarity search to RNA secondary structure prediction, to search for putative snRNA genes (including their promoters) and to examine their pattern of conservation among 11 available insect genomes (A. mellifera, Tribolium castaneum, Bombyx mori, Anopheles gambiae, Aedes aegypti, and six Drosophila species). We identified candidates for all nine spliceosomal snRNA genes in all the analyzed genomes. All the species contain a similar number of snRNA genes, with the exception of A. aegypti, whose genome contains more U1, U2, and U5 genes, and A. mellifera, whose genome contains fewer U2 and U5 genes. We found that snRNA genes are generally more closely related to homologs within the same genus than to those in other genera. Promoter regions for all spliceosomal snRNA genes within each insect species share similar sequence motifs that are likely to correspond to the PSEA (proximal sequence element A), the binding site for snRNA activating protein complex, but these promoter elements vary in sequence among the five insect families surveyed here. In contrast to the other insect species investigated, Dipteran genomes are characterized by a rapid evolution (or loss) of components of the U12 spliceosome and a striking loss of U12-type introns.

}, keywords = {Animals, Base Sequence, Bees, Computational Biology, Diptera, Evolution, Molecular, Genes, Insect, Genome, Insect, Molecular Sequence Data, Nucleic Acid Conformation, Phylogeny, Promoter Regions, Genetic, RNA Splicing, RNA, Small Nuclear, Sequence Analysis, RNA, Spliceosomes}, issn = {1355-8382}, doi = {10.1261/rna.259207}, author = {Mount, Stephen M and Gotea, Valer and Lin, Chiao-Feng and Hernandez, Kristina and Makalowski, Wojciech} } @article {49679, title = {SplicePort--an interactive splice-site analysis tool.}, journal = {Nucleic Acids Res}, volume = {35}, year = {2007}, month = {2007 Jul}, pages = {W285-91}, abstract = {

SplicePort is a web-based tool for splice-site analysis that allows the user to make splice-site predictions for submitted sequences. In addition, the user can also browse the rich catalog of features that underlies these predictions, and which we have found capable of providing high classification accuracy on human splice sites. Feature selection is optimized for human splice sites, but the selected features are likely to be predictive for other mammals as well. With our interactive feature browsing and visualization tool, the user can view and explore subsets of features used in splice-site prediction (either the features that account for the classification of a specific input sequence or the complete collection of features). Selected feature sets can be searched, ranked or displayed easily. The user can group features into clusters and frequency plot WebLogos can be generated for each cluster. The user can browse the identified clusters and their contributing elements, looking for new interesting signals, or can validate previously observed signals. The SplicePort web server can be accessed at http://www.cs.umd.edu/projects/SplicePort and http://www.spliceport.org.

}, keywords = {Base Sequence, Chromosome mapping, Computational Biology, Computer simulation, DNA, Genome, HUMANS, Internet, Models, Genetic, Molecular Sequence Data, Pattern Recognition, Automated, RNA Splice Sites, sequence alignment, Sequence Analysis, DNA, User-Computer Interface}, issn = {1362-4962}, doi = {10.1093/nar/gkm407}, author = {Dogan, Rezarta Islamaj and Getoor, Lise and Wilbur, W John and Mount, Stephen M} } @article {49641, title = {Analysis of fat body transcriptome from the adult tsetse fly, Glossina morsitans morsitans.}, journal = {Insect Mol Biol}, volume = {15}, year = {2006}, month = {2006 Aug}, pages = {411-24}, abstract = {

Tsetse flies (Diptera: Glossinidia) are vectors of pathogenic African trypanosomes. To develop a foundation for tsetse physiology, a normalized expressed sequence tag (EST) library was constructed from fat body tissue of immune-stimulated Glossina morsitans morsitans. Analysis of 20,257 high-quality ESTs yielded 6372 unique genes comprised of 3059 tentative consensus (TC) sequences and 3313 singletons (available at http://aksoylab.yale.edu). We analysed the putative fat body transcriptome based on homology to other gene products with known functions available in the public domain. In particular, we describe the immune-related products, reproductive function related yolk proteins and milk-gland protein, iron metabolism regulating ferritins and transferrin, and tsetse{\textquoteright}s major energy source proline biosynthesis. Expression analysis of the three yolk proteins indicates that all are detected in females, while only the yolk protein with similarity to lipases, is expressed in males. Milk gland protein, apparently important for larval nutrition, however, is primarily synthesized by accessory milk gland tissue.

}, keywords = {Adipose Tissue, Animals, Base Sequence, Computational Biology, DNA Primers, Egg Proteins, Expressed Sequence Tags, Female, Gene Expression Profiling, Insect Vectors, Male, Molecular Sequence Data, Reverse Transcriptase Polymerase Chain Reaction, Sequence Analysis, DNA, Sex Factors, Tsetse Flies}, issn = {0962-1075}, doi = {10.1111/j.1365-2583.2006.00649.x}, author = {Attardo, G M and Strickler-Dinglasan, P and Perkin, S A H and Caler, E and Bonaldo, M F and Soares, M B and El-Sayeed, N and Aksoy, S} } @article {38159, title = {Comparative genomic evidence for a close relationship between the dimorphic prosthecate bacteria Hyphomonas neptunium and Caulobacter crescentus}, journal = {Journal of bacteriologyJournal of bacteriology}, volume = {188}, year = {2006}, note = {http://www.ncbi.nlm.nih.gov/pubmed/16980487?dopt=Abstract}, type = {10.1128/JB.00111-06}, abstract = {The dimorphic prosthecate bacteria (DPB) are alpha-proteobacteria that reproduce in an asymmetric manner rather than by binary fission and are of interest as simple models of development. Prior to this work, the only member of this group for which genome sequence was available was the model freshwater organism Caulobacter crescentus. Here we describe the genome sequence of Hyphomonas neptunium, a marine member of the DPB that differs from C. crescentus in that H. neptunium uses its stalk as a reproductive structure. Genome analysis indicates that this organism shares more genes with C. crescentus than it does with Silicibacter pomeroyi (a closer relative according to 16S rRNA phylogeny), that it relies upon a heterotrophic strategy utilizing a wide range of substrates, that its cell cycle is likely to be regulated in a similar manner to that of C. crescentus, and that the outer membrane complements of H. neptunium and C. crescentus are remarkably similar. H. neptunium swarmer cells are highly motile via a single polar flagellum. With the exception of cheY and cheR, genes required for chemotaxis were absent in the H. neptunium genome. Consistent with this observation, H. neptunium swarmer cells did not respond to any chemotactic stimuli that were tested, which suggests that H. neptunium motility is a random dispersal mechanism for swarmer cells rather than a stimulus-controlled navigation system for locating specific environments. In addition to providing insights into bacterial development, the H. neptunium genome will provide an important resource for the study of other interesting biological processes including chromosome segregation, polar growth, and cell aging.}, keywords = {Alphaproteobacteria, Bacterial Outer Membrane Proteins, Caulobacter crescentus, cell cycle, Chemotaxis, DNA, Bacterial, Flagella, Genome, Bacterial, Microbial Viability, Molecular Sequence Data, Movement, Sequence Analysis, DNA, Sequence Homology, signal transduction}, author = {Badger, Jonathan H. and Hoover, Timothy R. and Brun, Yves V. and Weiner, Ronald M. and Laub, Michael T. and Alexandre, Gladys and Mr{\'a}zek, Jan and Ren, Qinghu and Paulsen, Ian T. and Nelson, Karen E. and Khouri, Hoda M. and Radune, Diana and Sosa, Julia and Dodson, Robert J. and Sullivan, Steven A. and Rosovitz, M. J. and Madupu, Ramana and Brinkac, Lauren M. and Durkin, A. Scott and Daugherty, Sean C. and Kothari, Sagar P. and Giglio, Michelle Gwinn and Zhou, Liwei and Haft, Daniel H. and J. Selengut and Davidsen, Tanja M. and Yang, Qi and Zafar, Nikhat and Ward, Naomi L.} } @article {38247, title = {Exopolysaccharide-associated protein sorting in environmental organisms: the PEP-CTERM/EpsH system. Application of a novel phylogenetic profiling heuristic}, journal = {BMC biologyBMC biology}, volume = {4}, year = {2006}, note = {http://www.ncbi.nlm.nih.gov/pubmed/16930487?dopt=Abstract}, type = {10.1186/1741-7007-4-29}, abstract = {BACKGROUND: Protein translocation to the proper cellular destination may be guided by various classes of sorting signals recognizable in the primary sequence. Detection in some genomes, but not others, may reveal sorting system components by comparison of the phylogenetic profile of the class of sorting signal to that of various protein families. RESULTS: We describe a short C-terminal homology domain, sporadically distributed in bacteria, with several key characteristics of protein sorting signals. The domain includes a near-invariant motif Pro-Glu-Pro (PEP). This possible recognition or processing site is followed by a predicted transmembrane helix and a cluster rich in basic amino acids. We designate this domain PEP-CTERM. It tends to occur multiple times in a genome if it occurs at all, with a median count of eight instances; Verrucomicrobium spinosum has sixty-five. PEP-CTERM-containing proteins generally contain an N-terminal signal peptide and exhibit high diversity and little homology to known proteins. All bacteria with PEP-CTERM have both an outer membrane and exopolysaccharide (EPS) production genes. By a simple heuristic for screening phylogenetic profiles in the absence of pre-formed protein families, we discovered that a homolog of the membrane protein EpsH (exopolysaccharide locus protein H) occurs in a species when PEP-CTERM domains are found. The EpsH family contains invariant residues consistent with a transpeptidase function. Most PEP-CTERM proteins are encoded by single-gene operons preceded by large intergenic regions. In the Proteobacteria, most of these upstream regions share a DNA sequence, a probable cis-regulatory site that contains a sigma-54 binding motif. The phylogenetic profile for this DNA sequence exactly matches that of three proteins: a sigma-54-interacting response regulator (PrsR), a transmembrane histidine kinase (PrsK), and a TPR protein (PrsT). CONCLUSION: These findings are consistent with the hypothesis that PEP-CTERM and EpsH form a protein export sorting system, analogous to the LPXTG/sortase system of Gram-positive bacteria, and correlated to EPS expression. It occurs preferentially in bacteria from sediments, soils, and biofilms. The novel method that led to these findings, partial phylogenetic profiling, requires neither global sequence clustering nor arbitrary similarity cutoffs and appears to be a rapid, effective alternative to other profiling methods.}, keywords = {Amino Acid Motifs, Amino Acid Sequence, bacteria, Bacterial Proteins, Biofilms, Genome, Bacterial, Markov chains, Molecular Sequence Data, Phylogeny, Polysaccharides, Bacterial, Protein Sorting Signals, Protein Transport, Seawater, sequence alignment, Soil Microbiology}, author = {Haft, Daniel H. and Paulsen, Ian T. and Ward, Naomi and J. Selengut} } @article {49640, title = {Trypanosoma cruzi mitochondrial maxicircles display species- and strain-specific variation and a conserved element in the non-coding region.}, journal = {BMC Genomics}, volume = {7}, year = {2006}, month = {2006}, pages = {60}, abstract = {

BACKGROUND: The mitochondrial DNA of kinetoplastid flagellates is distinctive in the eukaryotic world due to its massive size, complex form and large sequence content. Comprised of catenated maxicircles that contain rRNA and protein-coding genes and thousands of heterogeneous minicircles encoding small guide RNAs, the kinetoplast network has evolved along with an extreme form of mRNA processing in the form of uridine insertion and deletion RNA editing. Many maxicircle-encoded mRNAs cannot be translated without this post-transcriptional sequence modification.

RESULTS: We present the complete sequence and annotation of the Trypanosoma cruzi maxicircles for the CL Brener and Esmeraldo strains. Gene order is syntenic with Trypanosoma brucei and Leishmania tarentolae maxicircles. The non-coding components have strain-specific repetitive regions and a variable region that is unique for each strain with the exception of a conserved sequence element that may serve as an origin of replication, but shows no sequence identity with L. tarentolae or T. brucei. Alternative assemblies of the variable region demonstrate intra-strain heterogeneity of the maxicircle population. The extent of mRNA editing required for particular genes approximates that seen in T. brucei. Extensively edited genes were more divergent among the genera than non-edited and rRNA genes. Esmeraldo contains a unique 236-bp deletion that removes the 5{\textquoteright}-ends of ND4 and CR4 and the intergenic region. Esmeraldo shows additional insertions and deletions outside of areas edited in other species in ND5, MURF1, and MURF2, while CL Brener has a distinct insertion in MURF2.

CONCLUSION: The CL Brener and Esmeraldo maxicircles represent two of three previously defined maxicircle clades and promise utility as taxonomic markers. Restoration of the disrupted reading frames might be accomplished by strain-specific RNA editing. Elements in the non-coding region may be important for replication, transcription, and anchoring of the maxicircle within the kinetoplast network.

}, keywords = {Amino Acid Sequence, Animals, Animals, Inbred Strains, Base Composition, Conserved Sequence, DNA, Kinetoplast, Frameshifting, Ribosomal, Gene Deletion, Gene Order, Genetic Variation, Leishmania, Models, Biological, Molecular Sequence Data, Muscle Proteins, NADH Dehydrogenase, Open Reading Frames, Regulatory Elements, Transcriptional, RNA Editing, Sequence Homology, Amino Acid, Species Specificity, Trypanosoma brucei brucei, Trypanosoma cruzi, Ubiquitin-Protein Ligases, Untranslated Regions}, issn = {1471-2164}, doi = {10.1186/1471-2164-7-60}, author = {Westenberger, Scott J and Cerqueira, Gustavo C and El-Sayed, Najib M and Zingales, Bianca and Campbell, David A and Sturm, Nancy R} } @article {38287, title = {Genome analysis of multiple pathogenic isolates of Streptococcus agalactiae: implications for the microbial "pan-genome"}, journal = {Proceedings of the National Academy of Sciences of the United States of AmericaProceedings of the National Academy of Sciences of the United States of America}, volume = {102}, year = {2005}, note = {http://www.ncbi.nlm.nih.gov/pubmed/16172379?dopt=Abstract}, type = {10.1073/pnas.0506758102}, abstract = {The development of efficient and inexpensive genome sequencing methods has revolutionized the study of human bacterial pathogens and improved vaccine design. Unfortunately, the sequence of a single genome does not reflect how genetic variability drives pathogenesis within a bacterial species and also limits genome-wide screens for vaccine candidates or for antimicrobial targets. We have generated the genomic sequence of six strains representing the five major disease-causing serotypes of Streptococcus agalactiae, the main cause of neonatal infection in humans. Analysis of these genomes and those available in databases showed that the S. agalactiae species can be described by a pan-genome consisting of a core genome shared by all isolates, accounting for approximately 80\% of any single genome, plus a dispensable genome consisting of partially shared and strain-specific genes. Mathematical extrapolation of the data suggests that the gene reservoir available for inclusion in the S. agalactiae pan-genome is vast and that unique genes will continue to be identified even after sequencing hundreds of genomes.}, keywords = {Amino Acid Sequence, Bacterial Capsules, Base Sequence, Gene expression, Genes, Bacterial, Genetic Variation, Genome, Bacterial, Molecular Sequence Data, Phylogeny, sequence alignment, Sequence Analysis, DNA, Streptococcus agalactiae, virulence}, author = {Tettelin, Herv{\'e} and Masignani, Vega and Cieslewicz, Michael J. and Donati, Claudio and Medini, Duccio and Ward, Naomi L. and Angiuoli, Samuel V. and Crabtree, Jonathan and Jones, Amanda L. and Durkin, A. Scott and DeBoy, Robert T. and Davidsen, Tanja M. and Mora, Marirosa and Scarselli, Maria and Margarit y Ros, Immaculada and Peterson, Jeremy D. and Hauser, Christopher R. and Sundaram, Jaideep P. and Nelson, William C. and Madupu, Ramana and Brinkac, Lauren M. and Dodson, Robert J. and Rosovitz, Mary J. and Sullivan, Steven A. and Daugherty, Sean C. and Haft, Daniel H. and J. Selengut and Gwinn, Michelle L. and Zhou, Liwei and Zafar, Nikhat and Khouri, Hoda and Radune, Diana and Dimitrov, George and Watkins, Kisha and O{\textquoteright}Connor, Kevin J. B. and Smith, Shannon and Utterback, Teresa R. and White, Owen and Rubens, Craig E. and Grandi, Guido and Madoff, Lawrence C. and Kasper, Dennis L. and Telford, John L. and Wessels, Michael R. and Rappuoli, Rino and Fraser, Claire M.} } @article {49636, title = {Telomere and subtelomere of Trypanosoma cruzi chromosomes are enriched in (pseudo)genes of retrotransposon hot spot and trans-sialidase-like gene families: the origins of T. cruzi telomeres.}, journal = {Gene}, volume = {346}, year = {2005}, month = {2005 Feb 14}, pages = {153-61}, abstract = {

Here, we sequenced two large telomeric regions obtained from the pathogen protozoan Trypanosoma cruzi. These sequences, together with in silico assembled contigs, allowed us to establish the general features of telomeres and subtelomeres of this parasite. Our findings can be summarized as follows: We confirmed the presence of two types of telomeric ends; subtelomeric regions appeared to be enriched in (pseudo)genes of RHS (retrotransposon hot spot), TS (trans-sialidase)-like proteins, and putative surface protein DGF-1 (dispersed gene family-1). Sequence analysis of the ts-like genes located at the telomeres suggested that T. cruzi chromosomal ends could have been the site for generation of new gp85 variants, an important adhesin molecule involved in the invasion of mammalian cells by T. cruzi. Finally, a mechanism for generation of T. cruzi telomere by chromosome breakage and telomere healing is proposed.

}, keywords = {Amino Acid Sequence, Animals, Base Sequence, Chromosomes, Chromosomes, Artificial, Bacterial, DNA, Protozoan, Genes, Protozoan, Glycoproteins, Molecular Sequence Data, Multigene Family, Neuraminidase, Pseudogenes, Retroelements, Sequence Homology, Amino Acid, Sequence Homology, Nucleic Acid, Telomere, Trypanosoma cruzi}, issn = {0378-1119}, doi = {10.1016/j.gene.2004.10.014}, author = {Kim, Dong and Chiurillo, Miguel Angel and El-Sayed, Najib and Jones, Kristin and Santos, M{\'a}rcia R M and Porcile, Patricio E and Andersson, Bj{\"o}rn and Myler, Peter and da Silveira, Jose Franco and Ram{\'\i}rez, Jos{\'e} Luis} } @article {38575, title = {Whole-genome sequence analysis of Pseudomonas syringae pv. phaseolicola 1448A reveals divergence among pathovars in genes involved in virulence and transposition}, journal = {Journal of bacteriologyJournal of bacteriology}, volume = {187}, year = {2005}, note = {http://www.ncbi.nlm.nih.gov/pubmed/16159782?dopt=Abstract}, type = {10.1128/JB.187.18.6488-6498.2005}, abstract = {Pseudomonas syringae pv. phaseolicola, a gram-negative bacterial plant pathogen, is the causal agent of halo blight of bean. In this study, we report on the genome sequence of P. syringae pv. phaseolicola isolate 1448A, which encodes 5,353 open reading frames (ORFs) on one circular chromosome (5,928,787 bp) and two plasmids (131,950 bp and 51,711 bp). Comparative analyses with a phylogenetically divergent pathovar, P. syringae pv. tomato DC3000, revealed a strong degree of conservation at the gene and genome levels. In total, 4,133 ORFs were identified as putative orthologs in these two pathovars using a reciprocal best-hit method, with 3,941 ORFs present in conserved, syntenic blocks. Although these two pathovars are highly similar at the physiological level, they have distinct host ranges; 1448A causes disease in beans, and DC3000 is pathogenic on tomato and Arabidopsis. Examination of the complement of ORFs encoding virulence, fitness, and survival factors revealed a substantial, but not complete, overlap between these two pathovars. Another distinguishing feature between the two pathovars is their distinctive sets of transposable elements. With access to a fifth complete pseudomonad genome sequence, we were able to identify 3,567 ORFs that likely comprise the core Pseudomonas genome and 365 ORFs that are P. syringae specific.}, keywords = {Bacterial Proteins, DNA, Bacterial, Genes, Bacterial, Genome, Bacterial, Molecular Sequence Data, Pseudomonas syringae, Species Specificity, virulence}, author = {Joardar, Vinita and Lindeberg, Magdalen and Jackson, Robert W. and J. Selengut and Dodson, Robert and Brinkac, Lauren M. and Daugherty, Sean C. and Deboy, Robert and Durkin, A. Scott and Giglio, Michelle Gwinn and Madupu, Ramana and Nelson, William C. and Rosovitz, M. J. and Sullivan, Steven and Crabtree, Jonathan and Creasy, Todd and Davidsen, Tanja and Haft, Dan H. and Zafar, Nikhat and Zhou, Liwei and Halpin, Rebecca and Holley, Tara and Khouri, Hoda and Feldblyum, Tamara and White, Owen and Fraser, Claire M. and Chatterjee, Arun K. and Cartinhour, Sam and Schneider, David J. and Mansfield, John and Collmer, Alan and Buell, C. Robin} } @article {38165, title = {Comparison of the genome of the oral pathogen Treponema denticola with other spirochete genomes}, journal = {Proceedings of the National Academy of Sciences of the United States of AmericaProceedings of the National Academy of Sciences of the United States of America}, volume = {101}, year = {2004}, note = {http://www.ncbi.nlm.nih.gov/pubmed/15064399?dopt=Abstract}, type = {10.1073/pnas.0307639101}, abstract = {We present the complete 2,843,201-bp genome sequence of Treponema denticola (ATCC 35405) an oral spirochete associated with periodontal disease. Analysis of the T. denticola genome reveals factors mediating coaggregation, cell signaling, stress protection, and other competitive and cooperative measures, consistent with its pathogenic nature and lifestyle within the mixed-species environment of subgingival dental plaque. Comparisons with previously sequenced spirochete genomes revealed specific factors contributing to differences and similarities in spirochete physiology as well as pathogenic potential. The T. denticola genome is considerably larger in size than the genome of the related syphilis-causing spirochete Treponema pallidum. The differences in gene content appear to be attributable to a combination of three phenomena: genome reduction, lineage-specific expansions, and horizontal gene transfer. Genes lost due to reductive evolution appear to be largely involved in metabolism and transport, whereas some of the genes that have arisen due to lineage-specific expansions are implicated in various pathogenic interactions, and genes acquired via horizontal gene transfer are largely phage-related or of unknown function.}, keywords = {ATP-Binding Cassette Transporters, Bacterial Proteins, Base Sequence, Borrelia burgdorferi, Genes, Bacterial, Genome, Bacterial, Leptospira interrogans, Models, Genetic, Molecular Sequence Data, Mouth, Sequence Homology, Amino Acid, Treponema, Treponema pallidum}, author = {Seshadri, Rekha and Myers, Garry S. A. and Tettelin, Herv{\'e} and Eisen, Jonathan A. and Heidelberg, John F. and Dodson, Robert J. and Davidsen, Tanja M. and DeBoy, Robert T. and Fouts, Derrick E. and Haft, Dan H. and J. Selengut and Ren, Qinghu and Brinkac, Lauren M. and Madupu, Ramana and Kolonay, Jamie and Durkin, A. Scott and Daugherty, Sean C. and Shetty, Jyoti and Shvartsbeyn, Alla and Gebregeorgis, Elizabeth and Geer, Keita and Tsegaye, Getahun and Malek, Joel and Ayodeji, Bola and Shatsman, Sofiya and McLeod, Michael P. and Smajs, David and Howell, Jerrilyn K. and Pal, Sangita and Amin, Anita and Vashisth, Pankaj and McNeill, Thomas Z. and Xiang, Qin and Sodergren, Erica and Baca, Ernesto and Weinstock, George M. and Norris, Steven J. and Fraser, Claire M. and Paulsen, Ian T.} } @article {49684, title = {The Drosophila U1-70K protein is required for viability, but its arginine-rich domain is dispensable.}, journal = {Genetics}, volume = {168}, year = {2004}, month = {2004 Dec}, pages = {2059-65}, abstract = {

The conserved spliceosomal U1-70K protein is thought to play a key role in RNA splicing by linking the U1 snRNP particle to regulatory RNA-binding proteins. Although these protein interactions are mediated by repeating units rich in arginines and serines (RS domains) in vitro, tests of this domain{\textquoteright}s importance in intact multicellular organisms have not been carried out. Here we report a comprehensive genetic analysis of U1-70K function in Drosophila. Consistent with the idea that U1-70K is an essential splicing factor, we find that loss of U1-70K function results in lethality during embryogenesis. Surprisingly, and contrary to the current view of U1-70K function, animals carrying a mutant U1-70K protein lacking the arginine-rich domain, which includes two embedded sets of RS dipeptide repeats, have no discernible mutant phenotype. Through double-mutant studies, however, we show that the U1-70K RS domain deletion no longer supports viability when combined with a viable mutation in another U1 snRNP component. Together our studies demonstrate that while the protein interactions mediated by the U1-70K RS domain are not essential for viability, they nevertheless contribute to an essential U1 snRNP function.

}, keywords = {Amino Acid Sequence, Animals, Animals, Genetically Modified, Arginine, Drosophila, Drosophila Proteins, Molecular Sequence Data, Mutation, Protein Structure, Tertiary, Ribonucleoprotein, U1 Small Nuclear, RNA-Binding Proteins}, issn = {0016-6731}, doi = {10.1534/genetics.104.032532}, author = {Salz, Helen K and Mancebo, Ricardo S Y and Nagengast, Alexis A and Speck, Olga and Psotka, Mitchell and Mount, Stephen M} } @article {38302, title = {Genome sequence of Silicibacter pomeroyi reveals adaptations to the marine environment}, journal = {NatureNature}, volume = {432}, year = {2004}, note = {http://www.ncbi.nlm.nih.gov/pubmed/15602564?dopt=Abstract}, type = {10.1038/nature03170}, abstract = {Since the recognition of prokaryotes as essential components of the oceanic food web, bacterioplankton have been acknowledged as catalysts of most major biogeochemical processes in the sea. Studying heterotrophic bacterioplankton has been challenging, however, as most major clades have never been cultured or have only been grown to low densities in sea water. Here we describe the genome sequence of Silicibacter pomeroyi, a member of the marine Roseobacter clade (Fig. 1), the relatives of which comprise approximately 10-20\% of coastal and oceanic mixed-layer bacterioplankton. This first genome sequence from any major heterotrophic clade consists of a chromosome (4,109,442 base pairs) and megaplasmid (491,611 base pairs). Genome analysis indicates that this organism relies upon a lithoheterotrophic strategy that uses inorganic compounds (carbon monoxide and sulphide) to supplement heterotrophy. Silicibacter pomeroyi also has genes advantageous for associations with plankton and suspended particles, including genes for uptake of algal-derived compounds, use of metabolites from reducing microzones, rapid growth and cell-density-dependent regulation. This bacterium has a physiology distinct from that of marine oligotrophs, adding a new strategy to the recognized repertoire for coping with a nutrient-poor ocean.}, keywords = {Adaptation, Physiological, Carrier Proteins, Genes, Bacterial, Genome, Bacterial, marine biology, Molecular Sequence Data, Oceans and Seas, Phylogeny, plankton, RNA, Ribosomal, 16S, Roseobacter, Seawater}, author = {Moran, Mary Ann and Buchan, Alison and Gonz{\'a}lez, Jos{\'e} M. and Heidelberg, John F. and Whitman, William B. and Kiene, Ronald P. and Henriksen, James R. and King, Gary M. and Belas, Robert and Fuqua, Clay and Brinkac, Lauren and Lewis, Matt and Johri, Shivani and Weaver, Bruce and Pai, Grace and Eisen, Jonathan A. and Rahe, Elisha and Sheldon, Wade M. and Ye, Wenying and Miller, Todd R. and Carlton, Jane and Rasko, David A. and Paulsen, Ian T. and Ren, Qinghu and Daugherty, Sean C. and DeBoy, Robert T. and Dodson, Robert J. and Durkin, A. Scott and Madupu, Ramana and Nelson, William C. and Sullivan, Steven A. and Rosovitz, M. J. and Haft, Daniel H. and J. Selengut and Ward, Naomi} } @article {38303, title = {The genome sequence of the anaerobic, sulfate-reducing bacterium Desulfovibrio vulgaris Hildenborough}, journal = {Nature biotechnologyNature biotechnology}, volume = {22}, year = {2004}, note = {http://www.ncbi.nlm.nih.gov/pubmed/15077118?dopt=Abstract}, type = {10.1038/nbt959}, abstract = {Desulfovibrio vulgaris Hildenborough is a model organism for studying the energy metabolism of sulfate-reducing bacteria (SRB) and for understanding the economic impacts of SRB, including biocorrosion of metal infrastructure and bioremediation of toxic metal ions. The 3,570,858 base pair (bp) genome sequence reveals a network of novel c-type cytochromes, connecting multiple periplasmic hydrogenases and formate dehydrogenases, as a key feature of its energy metabolism. The relative arrangement of genes encoding enzymes for energy transduction, together with inferred cellular location of the enzymes, provides a basis for proposing an expansion to the {\textquoteright}hydrogen-cycling{\textquoteright} model for increasing energy efficiency in this bacterium. Plasmid-encoded functions include modification of cell surface components, nitrogen fixation and a type-III protein secretion system. This genome sequence represents a substantial step toward the elucidation of pathways for reduction (and bioremediation) of pollutants such as uranium and chromium and offers a new starting point for defining this organism{\textquoteright}s complex anaerobic respiration.}, keywords = {Desulfovibrio vulgaris, Energy Metabolism, Genome, Bacterial, Molecular Sequence Data}, author = {Heidelberg, John F. and Seshadri, Rekha and Haveman, Shelley A. and Hemme, Christopher L. and Paulsen, Ian T. and Kolonay, James F. and Eisen, Jonathan A. and Ward, Naomi and Methe, Barbara and Brinkac, Lauren M. and Daugherty, Sean C. and DeBoy, Robert T. and Dodson, Robert J. and Durkin, A. Scott and Madupu, Ramana and Nelson, William C. and Sullivan, Steven A. and Fouts, Derrick and Haft, Daniel H. and J. Selengut and Peterson, Jeremy D. and Davidsen, Tanja M. and Zafar, Nikhat and Zhou, Liwei and Radune, Diana and Dimitrov, George and Hance, Mark and Tran, Kevin and Khouri, Hoda and Gill, John and Utterback, Terry R. and Feldblyum, Tamara V. and Wall, Judy D. and Voordouw, Gerrit and Fraser, Claire M.} } @article {49634, title = {The ingi and RIME non-LTR retrotransposons are not randomly distributed in the genome of Trypanosoma brucei.}, journal = {Mol Biol Evol}, volume = {21}, year = {2004}, month = {2004 Mar}, pages = {520-8}, abstract = {

The ingi (long and autonomous) and RIME (short and nonautonomous) non--long-terminal repeat retrotransposons are the most abundant mobile elements characterized to date in the genome of the African trypanosome Trypanosoma brucei. These retrotransposons were thought to be randomly distributed, but a detailed and comprehensive analysis of their genomic distribution had not been performed until now. To address this question, we analyzed the ingi/RIME sequences and flanking sequences from the ongoing T. brucei genome sequencing project (TREU927/4 strain). Among the 81 ingi/RIME elements analyzed, 60\% are complete, and 7\% of the ingi elements (approximately 15 copies per haploid genome) appear to encode for their own transposition. The size of the direct repeat flanking the ingi/RIME retrotransposons is conserved (i.e., 12-bp), and a strong 11-bp consensus pattern precedes the 5{\textquoteright}-direct repeat. The presence of a consensus pattern upstream of the retroelements was confirmed by the analysis of the base occurrence in 294 GSS containing 5{\textquoteright}-adjacent ingi/RIME sequences. The conserved sequence is present upstream of ingis and RIMEs, suggesting that ingi-encoded enzymatic activities are used for retrotransposition of RIMEs, which are short nonautonomous retroelements. In conclusion, the ingi and RIME retroelements are not randomly distributed in the genome of T. brucei and are preceded by a conserved sequence, which may be the recognition site of the ingi-encoded endonuclease.

}, keywords = {Amino Acid Sequence, Animals, Base Sequence, Consensus Sequence, Genome, Protozoan, Molecular Sequence Data, Retroelements, Sequence Analysis, Trypanosoma brucei brucei}, issn = {0737-4038}, doi = {10.1093/molbev/msh045}, author = {Bringaud, Frederic and Biteau, Nicolas and Zuiderwijk, Eduard and Berriman, Matthew and El-Sayed, Najib M and Ghedin, Elodie and Melville, Sara E and Hall, Neil and Baltz, Th{\'e}o} } @article {38514, title = {Structural flexibility in the Burkholderia mallei genome}, journal = {Proceedings of the National Academy of Sciences of the United States of AmericaProceedings of the National Academy of Sciences of the United States of America}, volume = {101}, year = {2004}, note = {http://www.ncbi.nlm.nih.gov/pubmed/15377793?dopt=Abstract}, type = {10.1073/pnas.0403306101}, abstract = {The complete genome sequence of Burkholderia mallei ATCC 23344 provides insight into this highly infectious bacterium{\textquoteright}s pathogenicity and evolutionary history. B. mallei, the etiologic agent of glanders, has come under renewed scientific investigation as a result of recent concerns about its past and potential future use as a biological weapon. Genome analysis identified a number of putative virulence factors whose function was supported by comparative genome hybridization and expression profiling of the bacterium in hamster liver in vivo. The genome contains numerous insertion sequence elements that have mediated extensive deletions and rearrangements of the genome relative to Burkholderia pseudomallei. The genome also contains a vast number (>12,000) of simple sequence repeats. Variation in simple sequence repeats in key genes can provide a mechanism for generating antigenic variation that may account for the mammalian host{\textquoteright}s inability to mount a durable adaptive immune response to a B. mallei infection.}, keywords = {Animals, Base Composition, Base Sequence, Burkholderia mallei, Chromosomes, Bacterial, Cricetinae, Genome, Bacterial, Glanders, Liver, Mesocricetus, Molecular Sequence Data, Multigene Family, Oligonucleotide Array Sequence Analysis, Open Reading Frames, virulence}, author = {Nierman, William C. and DeShazer, David and Kim, H. Stanley and Tettelin, Herv{\'e} and Nelson, Karen E. and Feldblyum, Tamara and Ulrich, Ricky L. and Ronning, Catherine M. and Brinkac, Lauren M. and Daugherty, Sean C. and Davidsen, Tanja D. and DeBoy, Robert T. and Dimitrov, George and Dodson, Robert J. and Durkin, A. Scott and Gwinn, Michelle L. and Haft, Daniel H. and Khouri, Hoda and Kolonay, James F. and Madupu, Ramana and Mohammoud, Yasmin and Nelson, William C. and Radune, Diana and Romero, Claudia M. and Sarria, Saul and J. Selengut and Shamblin, Christine and Sullivan, Steven A. and White, Owen and Yu, Yan and Zafar, Nikhat and Zhou, Liwei and Fraser, Claire M.} } @article {38578, title = {X-ray crystal structure of the hypothetical phosphotyrosine phosphatase MDP-1 of the haloacid dehalogenase superfamily}, journal = {BiochemistryBiochemistry}, volume = {43}, year = {2004}, note = {http://www.ncbi.nlm.nih.gov/pubmed/15461449?dopt=Abstract}, type = {10.1021/bi0490688}, abstract = {The haloacid dehalogenase (HAD) superfamily is comprised of structurally homologous enzymes that share several conserved sequence motifs (loops I-IV) in their active site. The majority of HAD members are phosphohydrolases and may be divided into three subclasses depending on domain organization. In classes I and II, a mobile "cap" domain reorients upon substrate binding, closing the active site to bulk solvent. Members of the third class lack this additional domain. Herein, we report the 1.9 A X-ray crystal structures of a member of the third subclass, magnesium-dependent phosphatase-1 (MDP-1) both in its unliganded form and with the product analogue, tungstate, bound to the active site. The secondary structure of MDP-1 is similar to that of the "core" domain of other type I and type II HAD members with the addition of a small, 28-amino acid insert that does not close down to exclude bulk solvent in the presence of ligand. In addition, the monomeric oligomeric state of MDP-1 does not allow the participation of a second subunit in the formation and solvent protection of the active site. The binding sites for the phosphate portion of the substrate and Mg(II) cofactor are also similar to those of other HAD members, with all previously observed contacts conserved. Unlike other subclass III HAD members, MDP-1 appears to be equally able to dephosphorylate phosphotyrosine and closed-ring phosphosugars. Modeling of possible substrates in the active site of MDP-1 reveals very few potential interactions with the substrate leaving group. The mapping of conserved residues in sequences of MDP-1 from different eukaryotic organisms reveals that they colocalize to a large region on the surface of the protein outside the active site. This observation combined with the modeling studies suggests that the target of MDP-1 is most likely a phosphotyrosine in an unknown protein rather than a small sugar-based substrate.}, keywords = {Amino Acid Sequence, Animals, Binding Sites, Crystallography, X-Ray, HUMANS, Hydrogen-Ion Concentration, Hydrolases, Magnesium, Mice, Models, Molecular, Molecular Sequence Data, Phosphoprotein Phosphatases, Phosphotyrosine, Protein Phosphatase 1, Protein Structure, Quaternary, Protein Structure, Tertiary, sequence alignment, Solvents, Substrate Specificity}, author = {Peisach, Ezra and J. Selengut and Dunaway-Mariano, Debra and Allen, Karen N.} } @article {38168, title = {The complete genome sequence of the Arabidopsis and tomato pathogen Pseudomonas syringae pv. tomato DC3000}, journal = {Proceedings of the National Academy of Sciences of the United States of AmericaProceedings of the National Academy of Sciences of the United States of America}, volume = {100}, year = {2003}, note = {http://www.ncbi.nlm.nih.gov/pubmed/12928499?dopt=Abstract}, type = {10.1073/pnas.1731982100}, abstract = {We report the complete genome sequence of the model bacterial pathogen Pseudomonas syringae pathovar tomato DC3000 (DC3000), which is pathogenic on tomato and Arabidopsis thaliana. The DC3000 genome (6.5 megabases) contains a circular chromosome and two plasmids, which collectively encode 5,763 ORFs. We identified 298 established and putative virulence genes, including several clusters of genes encoding 31 confirmed and 19 predicted type III secretion system effector proteins. Many of the virulence genes were members of paralogous families and also were proximal to mobile elements, which collectively comprise 7\% of the DC3000 genome. The bacterium possesses a large repertoire of transporters for the acquisition of nutrients, particularly sugars, as well as genes implicated in attachment to plant surfaces. Over 12\% of the genes are dedicated to regulation, which may reflect the need for rapid adaptation to the diverse environments encountered during epiphytic growth and pathogenesis. Comparative analyses confirmed a high degree of similarity with two sequenced pseudomonads, Pseudomonas putida and Pseudomonas aeruginosa, yet revealed 1,159 genes unique to DC3000, of which 811 lack a known function.}, keywords = {Arabidopsis, Base Sequence, Biological Transport, Genome, Bacterial, Lycopersicon esculentum, Molecular Sequence Data, Plant Growth Regulators, Plasmids, Pseudomonas, Reactive Oxygen Species, Siderophores, virulence}, author = {Buell, C. Robin and Joardar, Vinita and Lindeberg, Magdalen and J. Selengut and Paulsen, Ian T. and Gwinn, Michelle L. and Dodson, Robert J. and DeBoy, Robert T. and Durkin, A. Scott and Kolonay, James F. and Madupu, Ramana and Daugherty, Sean and Brinkac, Lauren and Beanan, Maureen J. and Haft, Daniel H. and Nelson, William C. and Davidsen, Tanja and Zafar, Nikhat and Zhou, Liwei and Liu, Jia and Yuan, Qiaoping and Khouri, Hoda and Fedorova, Nadia and Tran, Bao and Russell, Daniel and Berry, Kristi and Utterback, Teresa and Aken, Susan E. van and Feldblyum, Tamara V. and D{\textquoteright}Ascenzo, Mark and Deng, Wen-Ling and Ramos, Adela R. and Alfano, James R. and Cartinhour, Samuel and Chatterjee, Arun K. and Delaney, Terrence P. and Lazarowitz, Sondra G. and Martin, Gregory B. and Schneider, David J. and Tang, Xiaoyan and Bender, Carol L. and White, Owen and Fraser, Claire M. and Collmer, Alan} } @article {49633, title = {The sequence and analysis of Trypanosoma brucei chromosome II.}, journal = {Nucleic Acids Res}, volume = {31}, year = {2003}, month = {2003 Aug 15}, pages = {4856-63}, abstract = {

We report here the sequence of chromosome II from Trypanosoma brucei, the causative agent of African sleeping sickness. The 1.2-Mb pairs encode about 470 predicted genes organised in 17 directional clusters on either strand, the largest cluster of which has 92 genes lined up over a 284-kb region. An analysis of the GC skew reveals strand compositional asymmetries that coincide with the distribution of protein-coding genes, suggesting these asymmetries may be the result of transcription-coupled repair on coding versus non-coding strand. A 5-cM genetic map of the chromosome reveals recombinational {\textquoteright}hot{\textquoteright} and {\textquoteright}cold{\textquoteright} regions, the latter of which is predicted to include the putative centromere. One end of the chromosome consists of a 250-kb region almost exclusively composed of RHS (pseudo)genes that belong to a newly characterised multigene family containing a hot spot of insertion for retroelements. Interspersed with the RHS genes are a few copies of truncated RNA polymerase pseudogenes as well as expression site associated (pseudo)genes (ESAGs) 3 and 4, and 76 bp repeats. These features are reminiscent of a vestigial variant surface glycoprotein (VSG) gene expression site. The other end of the chromosome contains a 30-kb array of VSG genes, the majority of which are pseudogenes, suggesting that this region may be a site for modular de novo construction of VSG gene diversity during transposition/gene conversion events.

}, keywords = {Animals, Antigens, Protozoan, Chromosome mapping, Chromosomes, DNA, Protozoan, Gene Duplication, Genes, Protozoan, Molecular Sequence Data, Pseudogenes, Recombination, Genetic, Sequence Analysis, DNA, Trypanosoma brucei brucei}, issn = {1362-4962}, author = {el-Sayed, Najib M A and Ghedin, Elodie and Song, Jinming and MacLeod, Annette and Bringaud, Frederic and Larkin, Christopher and Wanless, David and Peterson, Jeremy and Hou, Lihua and Taylor, Sonya and Tweedie, Alison and Biteau, Nicolas and Khalak, Hanif G and Lin, Xiaoying and Mason, Tanya and Hannick, Linda and Caler, Elisabet and Blandin, Ga{\"e}lle and Bartholomeu, Daniella and Simpson, Anjana J and Kaul, Samir and Zhao, Hong and Pai, Grace and Van Aken, Susan and Utterback, Teresa and Haas, Brian and Koo, Hean L and Umayam, Lowell and Suh, Bernard and Gerrard, Caroline and Leech, Vanessa and Qi, Rong and Zhou, Shiguo and Schwartz, David and Feldblyum, Tamara and Salzberg, Steven and Tait, Andrew and Turner, C Michael R and Ullu, Elisabetta and White, Owen and Melville, Sara and Adams, Mark D and Fraser, Claire M and Donelson, John E} } @article {49686, title = {Sex-lethal splicing autoregulation in vivo: interactions between SEX-LETHAL, the U1 snRNP and U2AF underlie male exon skipping.}, journal = {Development}, volume = {130}, year = {2003}, month = {2003 Feb}, pages = {463-71}, abstract = {

Alternative splicing of the Sex-lethal pre-mRNA has long served as a model example of a regulated splicing event, yet the mechanism by which the female-specific SEX-LETHAL RNA-binding protein prevents inclusion of the translation-terminating male exon is not understood. Thus far, the only general splicing factor for which there is in vivo evidence for a regulatory role in the pathway leading to male-exon skipping is sans-fille (snf), a protein component of the spliceosomal U1 and U2 snRNPs. Its role, however, has remained enigmatic because of questions about whether SNF acts as part of an intact snRNP or a free protein. We provide evidence that SEX-LETHAL interacts with SANS-FILLE in the context of the U1 snRNP, through the characterization of a point mutation that interferes with both assembly into the U1 snRNP and complex formation with SEX-LETHAL. Moreover, we find that SEX-LETHAL associates with other integral U1 snRNP components, and we provide genetic evidence to support the biological relevance of these physical interactions. Similar genetic and biochemical approaches also link SEX-LETHAL with the heterodimeric splicing factor, U2AF. These studies point specifically to a mechanism by which SEX-LETHAL represses splicing by interacting with these key splicing factors at both ends of the regulated male exon. Moreover, because U2AF and the U1 snRNP are only associated transiently with the pre-mRNA during the course of spliceosome assembly, our studies are difficult to reconcile with the current model that proposes that the SEX-LETHAL blocks splicing at the second catalytic step, and instead argue that the SEX-LETHAL protein acts after splice site recognition, but before catalysis begins.

}, keywords = {Alternative Splicing, Amino Acid Sequence, Animals, Animals, Genetically Modified, Drosophila melanogaster, Drosophila Proteins, Exons, Female, Gene Expression Regulation, Developmental, Genes, Insect, Homeostasis, Male, Models, Genetic, Molecular Sequence Data, Nuclear Proteins, Point Mutation, Ribonucleoprotein, U1 Small Nuclear, Ribonucleoproteins, RNA Splicing, RNA-Binding Proteins, Sequence Homology, Amino Acid, Sex Differentiation}, issn = {0950-1991}, author = {Nagengast, Alexis A and Stitzinger, Shane M and Tseng, Chin-Hsiu and Mount, Stephen M and Salz, Helen K} } @article {38536, title = {The transcription factor Eyes absent is a protein tyrosine phosphatase}, journal = {NatureNature}, volume = {426}, year = {2003}, note = {http://www.ncbi.nlm.nih.gov/pubmed/14628053?dopt=Abstract}, type = {10.1038/nature02097}, abstract = {Post-translational modifications provide sensitive and flexible mechanisms to dynamically modulate protein function in response to specific signalling inputs. In the case of transcription factors, changes in phosphorylation state can influence protein stability, conformation, subcellular localization, cofactor interactions, transactivation potential and transcriptional output. Here we show that the evolutionarily conserved transcription factor Eyes absent (Eya) belongs to the phosphatase subgroup of the haloacid dehalogenase (HAD) superfamily, and propose a function for it as a non-thiol-based protein tyrosine phosphatase. Experiments performed in cultured Drosophila cells and in vitro indicate that Eyes absent has intrinsic protein tyrosine phosphatase activity and can autocatalytically dephosphorylate itself. Confirming the biological significance of this function, mutations that disrupt the phosphatase active site severely compromise the ability of Eyes absent to promote eye specification and development in Drosophila. Given the functional importance of phosphorylation-dependent modulation of transcription factor activity, this evidence for a nuclear transcriptional coactivator with intrinsic phosphatase activity suggests an unanticipated method of fine-tuning transcriptional regulation.}, keywords = {Amino Acid Motifs, Amino Acid Sequence, Animals, Antibodies, Phospho-Specific, Drosophila melanogaster, Drosophila Proteins, Embryonic Induction, eye, Eye Proteins, Gene Expression Regulation, Kinetics, Mice, Models, Molecular, Molecular Sequence Data, Mutation, Phosphorylation, Protein Conformation, Protein Tyrosine Phosphatases, Substrate Specificity, Transcription Factors}, author = {Tootle, Tina L. and Silver, Serena J. and Davies, Erin L. and Newman, Victoria and Latek, Robert R. and Mills, Ishara A. and J. Selengut and Parlikar, Beth E. W. and Rebay, Ilaria} } @article {49629, title = {Analysis of stage-specific gene expression in the bloodstream and the procyclic form of Trypanosoma brucei using a genomic DNA-microarray.}, journal = {Mol Biochem Parasitol}, volume = {123}, year = {2002}, month = {2002 Aug 28}, pages = {115-23}, abstract = {

A microarray comprising 21,024 different PCR products spotted on glass slides was constructed for gene expression studies on Trypanosoma brucei. The arrayed fragments were generated from a T. brucei shotgun clone library, which had been prepared from randomly sheared and size-fractionated genomic DNA. For the identification of stage-specific gene activity, total RNA from in vitro cultures of the human, long slender form and the insect, procyclic form of the parasite was labelled and hybridised to the microarray. Approximately 75\% of the genomic fragments produced a signal and about 2\% exhibited significant differences between the transcript levels in the bloodstream and procyclic forms. A few results were confirmed by Northern blot analysis or reverse-transcription and PCR. Three hundred differentially regulated clones have been selected for sequencing. So far, of 33 clones that showed about 2-fold or more over-expression in bloodstream forms, 15 contained sequences similar to those of VSG expression sites and at least six others appeared non-protein-coding. Of 29 procyclic-specific clones, at least eight appeared not to be protein-coding. A surprisingly large proportion of known regulated genes was already identified in this small sample, and some new ones were found, illustrating the utility of genomic arrays.

}, keywords = {Animals, Blotting, Northern, Escherichia coli, Gene expression, Gene Expression Profiling, Genes, Protozoan, HUMANS, Life Cycle Stages, Molecular Sequence Data, Oligonucleotide Array Sequence Analysis, Polymerase Chain Reaction, Transcription, Genetic, Trypanosoma brucei brucei}, issn = {0166-6851}, author = {Diehl, Susanne and Diehl, Frank and El-Sayed, Najib M and Clayton, Christine and Hoheisel, J{\"o}rg D} } @article {49687, title = {The draft genome of Ciona intestinalis: insights into chordate and vertebrate origins.}, journal = {Science}, volume = {298}, year = {2002}, month = {2002 Dec 13}, pages = {2157-67}, abstract = {

The first chordates appear in the fossil record at the time of the Cambrian explosion, nearly 550 million years ago. The modern ascidian tadpole represents a plausible approximation to these ancestral chordates. To illuminate the origins of chordate and vertebrates, we generated a draft of the protein-coding portion of the genome of the most studied ascidian, Ciona intestinalis. The Ciona genome contains approximately 16,000 protein-coding genes, similar to the number in other invertebrates, but only half that found in vertebrates. Vertebrate gene families are typically found in simplified form in Ciona, suggesting that ascidians contain the basic ancestral complement of genes involved in cell signaling and development. The ascidian genome has also acquired a number of lineage-specific innovations, including a group of genes engaged in cellulose metabolism that are related to those in bacteria and fungi.

}, keywords = {Alleles, Animals, Apoptosis, Base Sequence, Cellulose, Central Nervous System, Ciona intestinalis, Computational Biology, Endocrine System, Gene Dosage, Gene Duplication, genes, Genes, Homeobox, Genome, Heart, Immunity, Molecular Sequence Data, Multigene Family, Muscle Proteins, Organizers, Embryonic, Phylogeny, Polymorphism, Genetic, Proteins, Sequence Analysis, DNA, Sequence Homology, Nucleic Acid, Species Specificity, Thyroid Gland, Urochordata, Vertebrates}, issn = {1095-9203}, doi = {10.1126/science.1080049}, author = {Dehal, Paramvir and Satou, Yutaka and Campbell, Robert K and Chapman, Jarrod and Degnan, Bernard and De Tomaso, Anthony and Davidson, Brad and Di Gregorio, Anna and Gelpke, Maarten and Goodstein, David M and Harafuji, Naoe and Hastings, Kenneth E M and Ho, Isaac and Hotta, Kohji and Huang, Wayne and Kawashima, Takeshi and Lemaire, Patrick and Martinez, Diego and Meinertzhagen, Ian A and Necula, Simona and Nonaka, Masaru and Putnam, Nik and Rash, Sam and Saiga, Hidetoshi and Satake, Masanobu and Terry, Astrid and Yamada, Lixy and Wang, Hong-Gang and Awazu, Satoko and Azumi, Kaoru and Boore, Jeffrey and Branno, Margherita and Chin-Bow, Stephen and DeSantis, Rosaria and Doyle, Sharon and Francino, Pilar and Keys, David N and Haga, Shinobu and Hayashi, Hiroko and Hino, Kyosuke and Imai, Kaoru S and Inaba, Kazuo and Kano, Shungo and Kobayashi, Kenji and Kobayashi, Mari and Lee, Byung-In and Makabe, Kazuhiro W and Manohar, Chitra and Matassi, Giorgio and Medina, Monica and Mochizuki, Yasuaki and Mount, Steve and Morishita, Tomomi and Miura, Sachiko and Nakayama, Akie and Nishizaka, Satoko and Nomoto, Hisayo and Ohta, Fumiko and Oishi, Kazuko and Rigoutsos, Isidore and Sano, Masako and Sasaki, Akane and Sasakura, Yasunori and Shoguchi, Eiichi and Shin-i, Tadasu and Spagnuolo, Antoinetta and Stainier, Didier and Suzuki, Miho M and Tassy, Olivier and Takatori, Naohito and Tokuoka, Miki and Yagi, Kasumi and Yoshizaki, Fumiko and Wada, Shuichi and Zhang, Cindy and Hyatt, P Douglas and Larimer, Frank and Detter, Chris and Doggett, Norman and Glavina, Tijana and Hawkins, Trevor and Richardson, Paul and Lucas, Susan and Kohara, Yuji and Levine, Michael and Satoh, Nori and Rokhsar, Daniel S} } @article {49688, title = {Evidence for a plastid origin of plant ethylene receptor genes.}, journal = {Plant Physiol}, volume = {130}, year = {2002}, month = {2002 Sep}, pages = {10-4}, keywords = {Amino Acid Sequence, Anabaena, Arabidopsis, Cyanobacteria, Molecular Sequence Data, Plant Proteins, Plastids, Protein Kinases, Receptors, Cell Surface, Sequence Homology, Amino Acid}, issn = {0032-0889}, doi = {10.1104/pp.005397}, author = {Mount, Stephen M and Chang, Caren} } @article {38304, title = {Genome sequence of the human malaria parasite Plasmodium falciparum}, journal = {NatureNature}, volume = {419}, year = {2002}, note = {http://www.ncbi.nlm.nih.gov/pubmed/12368864?dopt=Abstract}, type = {10.1038/nature01097}, abstract = {The parasite Plasmodium falciparum is responsible for hundreds of millions of cases of malaria, and kills more than one million African children annually. Here we report an analysis of the genome sequence of P. falciparum clone 3D7. The 23-megabase nuclear genome consists of 14 chromosomes, encodes about 5,300 genes, and is the most (A + T)-rich genome sequenced to date. Genes involved in antigenic variation are concentrated in the subtelomeric regions of the chromosomes. Compared to the genomes of free-living eukaryotic microbes, the genome of this intracellular parasite encodes fewer enzymes and transporters, but a large proportion of genes are devoted to immune evasion and host-parasite interactions. Many nuclear-encoded proteins are targeted to the apicoplast, an organelle involved in fatty-acid and isoprenoid metabolism. The genome sequence provides the foundation for future studies of this organism, and is being exploited in the search for new drugs and vaccines to fight malaria.}, keywords = {Animals, Chromosome Structures, DNA Repair, DNA Replication, DNA, Protozoan, Evolution, Molecular, Genome, Protozoan, HUMANS, Malaria Vaccines, Malaria, Falciparum, Membrane Transport Proteins, Molecular Sequence Data, Plasmodium falciparum, Plastids, Proteome, Protozoan Proteins, Recombination, Genetic, Sequence Analysis, DNA}, author = {Gardner, Malcolm J. and Hall, Neil and Fung, Eula and White, Owen and Berriman, Matthew and Hyman, Richard W. and Carlton, Jane M. and Pain, Arnab and Nelson, Karen E. and Bowman, Sharen and Paulsen, Ian T. and James, Keith and Eisen, Jonathan A. and Rutherford, Kim and Salzberg, Steven L. and Craig, Alister and Kyes, Sue and Chan, Man-Suen and Nene, Vishvanath and Shallom, Shamira J. and Suh, Bernard and Peterson, Jeremy and Angiuoli, Sam and Pertea, Mihaela and Allen, Jonathan and J. Selengut and Haft, Daniel and Mather, Michael W. and Vaidya, Akhil B. and Martin, David M. A. and Fairlamb, Alan H. and Fraunholz, Martin J. and Roos, David S. and Ralph, Stuart A. and McFadden, Geoffrey I. and Cummings, Leda M. and Subramanian, G. Mani and Mungall, Chris and Venter, J. Craig and Carucci, Daniel J. and Hoffman, Stephen L. and Newbold, Chris and Davis, Ronald W. and Fraser, Claire M. and Barrell, Bart} } @article {49630, title = {Identification of non-autonomous non-LTR retrotransposons in the genome of Trypanosoma cruzi.}, journal = {Mol Biochem Parasitol}, volume = {124}, year = {2002}, month = {2002 Sep-Oct}, pages = {73-8}, abstract = {

As observed for most eukaryotic cells, trypanosomatids contains non-LTR retrotransposons randomly inserted in the nuclear genome. Autonomous retroelements which, code for their own transposition, have been characterized in Trypanosoma brucei (ingi) and Trypanosoma cruzi (L1Tc), whereas non-autonomous retroelements have only been characterized in T. brucei (RIME). Here, we have characterized in the genome of Trypanosoma cruzi four complete copies of a non-autonomous non-LTR retrotransposon, called NARTc. This 0.26 kb NARTc element has the characteristics of non-LTR retrotransposons: the presence a poly(dA) tail and of a short flanking duplicated motif. Analysis of the Genome Survey Sequence databases indicated that the Trypanosoma cruzi haploid genome contains about 140 NARTc copies and about twice as many L1Tc copies. Interestingly, the NARTc and L1Tc retroelements share, with the Trypanosoma brucei ingi and RIME retrotransposons, a common sequence (the first 45 bp with 91\% identity), whereas the remaining sequences are very divergent. This suggests that these four trypanosome non-LTR retrotransposons were derived from the same common ancester and the sequence of their 5{\textquoteright}-extremity may have a functional role. In addition, the genome of Leishmania major contains the same conserved motif present in the trypanosome retroelements, whicle no transposable elements have been detected so far in Leishmania sp.

}, keywords = {Animals, Base Sequence, Computational Biology, Genome, Protozoan, Long Interspersed Nucleotide Elements, Molecular Sequence Data, Retroelements, Short Interspersed Nucleotide Elements, Trypanosoma cruzi}, issn = {0166-6851}, author = {Bringaud, Frederic and Garc{\'\i}a-P{\'e}rez, Jos{\'e} Luis and Heras, Sara R and Ghedin, Elodie and El-Sayed, Najib M and Andersson, Bj{\"o}rn and Baltz, Th{\'e}o and Lopez, Manuel C} } @article {49631, title = {A new, expressed multigene family containing a hot spot for insertion of retroelements is associated with polymorphic subtelomeric regions of Trypanosoma brucei.}, journal = {Eukaryot Cell}, volume = {1}, year = {2002}, month = {2002 Feb}, pages = {137-51}, abstract = {

We describe a novel gene family that forms clusters in subtelomeric regions of Trypanosoma brucei chromosomes and partially accounts for the observed clustering of retrotransposons. The ingi and ribosomal inserted mobile element (RIME) non-LTR retrotransposons share 250 bp at both extremities and are the most abundant putatively mobile elements, with about 500 copies per haploid genome. From cDNA clones and subsequently in the T. brucei genomic DNA databases, we identified 52 homologous gene and pseudogene sequences, 16 of which contain a RIME and/or ingi retrotransposon inserted at exactly the same relative position. Here these genes are called the RHS family, for retrotransposon hot spot. Comparison of the protein sequences encoded by RHS genes (21 copies) and pseudogenes (24 copies) revealed a conserved central region containing an ATP/GTP-binding motif and the RIME/ingi insertion site. The RHS proteins share between 13 and 96\% identity, and six subfamilies, RHS1 to RHS6, can be defined on the basis of their divergent C-terminal domains. Immunofluorescence and Western blot analyses using RHS subfamily-specific immune sera show that RHS proteins are constitutively expressed and occur mainly in the nucleus. Analysis of Genome Survey Sequence databases indicated that the Trypanosoma brucei diploid genome contains about 280 RHS (pseudo)genes. Among the 52 identified RHS (pseudo)genes, 48 copies are in three RHS clusters located in subtelomeric regions of chromosomes Ia and II and adjacent to the active bloodstream form expression site in T. brucei strain TREU927/4 GUTat10.1. RHS genes comprise the remaining sequence of the size-polymorphic "repetitive region" described for T. brucei chromosome I, and a homologous gene family is present in the Trypanosoma cruzi genome.

}, keywords = {Amino Acid Sequence, Animals, Base Sequence, Cloning, Molecular, DNA Primers, DNA, Protozoan, Escherichia coli, Genes, Protozoan, Molecular Sequence Data, Multigene Family, Mutagenesis, Insertional, Phylogeny, Polymorphism, Genetic, Protozoan Proteins, Pseudogenes, Retroelements, sequence alignment, Sequence Homology, Amino Acid, Telomere, Trypanosoma brucei brucei, Trypanosoma cruzi}, issn = {1535-9778}, author = {Bringaud, Frederic and Biteau, Nicolas and Melville, Sara E and Hez, St{\'e}phanie and El-Sayed, Najib M and Leech, Vanessa and Berriman, Matthew and Hall, Neil and Donelson, John E and Baltz, Th{\'e}o} } @article {49632, title = {Trypanosoma cruzi: RNA structure and post-transcriptional control of tubulin gene expression.}, journal = {Exp Parasitol}, volume = {102}, year = {2002}, month = {2002 Nov-Dec}, pages = {123-33}, abstract = {

Changes in tubulin expression are among the biochemical and morphological adaptations that occur during the life cycle of Trypanosomatids. To investigate the mechanism responsible for the differential accumulation of tubulin mRNAs in Trypanosoma cruzi, we determine the sequences of alpha- and beta-tubulin transcripts and analyzed their expression during the life cycle of the parasite. Two beta-tubulin mRNAs of 1.9 and 2.3 kb were found to differ mainly by an additional 369 nucleotides at the end of the 3{\textquoteright} untranslated region (UTR). Although their transcription rates are similar in epimastigotes and amastigotes, alpha- and beta-tubulin transcripts are 3- to 6-fold more abundant in epimastigotes than in trypomastigotes and amastigotes. Accordingly, the half-lives of alpha- and beta-tubulin mRNAs are significantly higher in epimastigotes than in amastigotes. Transient transfection experiments indicated that positive regulatory elements occur in the 3{\textquoteright} UTR plus downstream intergenic region of the alpha-tubulin gene and that both positive and negative elements occur in the equivalent regions of the beta-tubulin gene.

}, keywords = {Animals, Base Sequence, Blotting, Northern, DNA, Complementary, DNA, Protozoan, Gene Expression Regulation, Half-Life, Life Cycle Stages, Molecular Sequence Data, RNA Processing, Post-Transcriptional, RNA, Messenger, RNA, Protozoan, Transcription, Genetic, Transfection, Trypanosoma cruzi, Tubulin}, issn = {0014-4894}, author = {Bartholomeu, Daniella C and Silva, Rosiane A and Galv{\~a}o, Lucia M C and el-Sayed, Najib M A and Donelson, John E and Teixeira, Santuza M R} } @article {38368, title = {MDP-1 is a new and distinct member of the haloacid dehalogenase family of aspartate-dependent phosphohydrolases}, journal = {BiochemistryBiochemistry}, volume = {40}, year = {2001}, note = {http://www.ncbi.nlm.nih.gov/pubmed/11601995?dopt=Abstract}, abstract = {MDP-1 is a eukaryotic magnesium-dependent acid phosphatase with little sequence homology to previously characterized phosphatases. The presence of a conserved motif (Asp-X-Asp-X-Thr) in the N terminus of MDP-1 suggested a relationship to the haloacid dehalogenase (HAD) superfamily, which contains a number of magnesium-dependent acid phosphatases. These phosphatases utilize an aspartate nucleophile and contain a number of conserved active-site residues and hydrophobic patches, which can be plausibly aligned with conserved residues in MDP-1. Seven site-specific point mutants of MDP-1 were produced by modifying the catalytic aspartate, serine, and lysine residues to asparagine or glutamate, alanine, and arginine, respectively. The activity of these mutants confirms the assignment of MDP-1 as a member of the HAD superfamily. Detailed comparison of the sequence of the 15 MDP-1 sequences from various organisms with other HAD superfamily sequences suggests that MDP-1 is not closely related to any particular member of the superfamily. The crystal structures of several HAD family enzymes identify a domain proximal to the active site responsible for important interactions with low molecular weight substrates. The absence of this domain or any other that might perform the same function in MDP-1 suggests an "open" active site capable of interactions with large substrates such as proteins. This suggestion was experimentally confirmed by demonstration that MDP-1 is competent to catalyze the dephosphorylation of tyrosine-phosphorylated proteins.}, keywords = {Amino Acid Motifs, Amino Acid Sequence, Animals, Aspartic Acid, Catalytic Domain, HUMANS, Hydrolases, Mice, Molecular Sequence Data, Multigene Family, Mutagenesis, Site-Directed, Phosphoprotein Phosphatases, Protein Structure, Tertiary, Protein Tyrosine Phosphatases, Rats, Saccharomyces cerevisiae, sequence alignment, Sequence Homology, Amino Acid}, author = {J. Selengut} } @article {38369, title = {MDP-1: A novel eukaryotic magnesium-dependent phosphatase}, journal = {BiochemistryBiochemistry}, volume = {39}, year = {2000}, note = {http://www.ncbi.nlm.nih.gov/pubmed/10889041?dopt=Abstract}, abstract = {We report here the purification, cloning, expression, and characterization of a novel phosphatase, MDP-1. In the course of investigating the reported acid phosphatase activity of carbonic anhydrase III preparations, several discrete phosphatases were discerned. One of these, a magnesium-dependent species of 18.6 kDa, was purified to homogeneity and yielded several peptide sequences from which the parent gene was identified by database searching. Although orthologous genes were identified in fungi and plants as well as mammalian species, there was no apparent homology to any known family of phosphatases. The enzyme was expressed in Escherichia coli with a fusion tag and purified by affinity methods. The recombinant enzyme showed magnesium-dependent acid phosphatase activity comparable to the originally isolated rabbit protein. The enzyme catalyzes the rapid hydrolysis of p-nitrophenyl phosphate, ribose-5-phosphate, and phosphotyrosine. The selectivity for phosphotyrosine over phosphoserine or phosphothreonine is considerable, but the enzyme did not show activity toward five phosphotyrosine-containing peptides. None of the various substrates assayed (including various nucleotide, sugar, amino acid and peptide phosphates, phosphoinositides, and phosphodiesters) exhibited K(M) values lower than 1 mM, and many showed negligible rates of hydrolysis. The enzyme is inhibited by vanadate and fluoride but not by azide, cyanide, calcium, lithium, or tartaric acid. Chemical labeling, refolding, dialysis, and mutagenesis experiments suggest that the enzymatic mechanism is not dependent on cysteine, histidine, or nonmagnesium metal ions. In recognition of these observations, the enzyme has been given the name magnesium-dependent phosphatase-1 (MDP-1).}, keywords = {Amino Acid Sequence, Animals, Catalysis, Cations, Chromatography, Affinity, Cloning, Molecular, Cysteine, Enzyme Inhibitors, Histidine, Hydrogen-Ion Concentration, Magnesium, Mice, Molecular Sequence Data, Phosphoprotein Phosphatases, Protein Phosphatase 1, Rabbits, Sequence Analysis, Protein, Sequence Homology, Amino Acid, Substrate Specificity}, author = {J. Selengut and Levine, R. L.} } @article {49693, title = {Genetic depletion reveals an essential role for an SR protein splicing factor in vertebrate cells.}, journal = {Bioessays}, volume = {19}, year = {1997}, month = {1997 Mar}, pages = {189-92}, abstract = {

SR proteins are essential for the splicing of messenger RNA precursors in vitro, where they also alter splice site selection in a concentration-dependent manner. Although experiments involving overexpression or dominant mutations have confirmed that these proteins can influence RNA processing decisions in vivo, similar results with loss-of-function mutations have been lacking. Now, a system for genetic depletion of the chicken B cell line DT40 has revealed that the SR protein ASF/SF2 (alternative splicing factor/splicing factor 2) is essential for viability in these cells(1). This study opens the way for a complete functional dissection of this protein, and other SR proteins, in vivo.

}, keywords = {Amino Acid Sequence, Animals, Molecular Sequence Data, Nuclear Proteins, RNA Splicing, RNA-Binding Proteins, Serine-Arginine Splicing Factors, Vertebrates}, issn = {0265-9247}, doi = {10.1002/bies.950190302}, author = {Mount, S M} } @article {49695, title = {AT-AC introns: an ATtACk on dogma.}, journal = {Science}, volume = {271}, year = {1996}, month = {1996 Mar 22}, pages = {1690-2}, keywords = {Animals, Base Composition, Base Sequence, Consensus Sequence, HUMANS, Introns, Molecular Sequence Data, Mutation, RNA Precursors, RNA Splicing, RNA, Small Nuclear, Spliceosomes}, issn = {0036-8075}, author = {Mount, S M} } @article {49696, title = {Genetic enhancement of RNA-processing defects by a dominant mutation in B52, the Drosophila gene for an SR protein splicing factor.}, journal = {Mol Cell Biol}, volume = {15}, year = {1995}, month = {1995 Nov}, pages = {6273-82}, abstract = {

SR proteins are essential for pre-mRNA splicing in vitro, act early in the splicing pathway, and can influence alternative splice site choice. Here we describe the isolation of both dominant and loss-of-function alleles of B52, the gene for a Drosophila SR protein. The allele B52ED was identified as a dominant second-site enhancer of white-apricot (wa), a retrotransposon insertion in the second intron of the eye pigmentation gene white with a complex RNA-processing defect. B52ED also exaggerates the mutant phenotype of a distinct white allele carrying a 5{\textquoteright} splice site mutation (wDR18), and alters the pattern of sex-specific splicing at doublesex under sensitized conditions, so that the male-specific splice is favored. In addition to being a dominant enhancer of these RNA-processing defects, B52ED is a recessive lethal allele that fails to complement other lethal alleles of B52. Comparison of B52ED with the B52+ allele from which it was derived revealed a single change in a conserved amino acid in the beta 4 strand of the first RNA-binding domain of B52, which suggests that altered RNA binding is responsible for the dominant phenotype. Reversion of the B52ED dominant allele with X rays led to the isolation of a B52 null allele. Together, these results indicate a critical role for the SR protein B52 in pre-mRNA splicing in vivo.

}, keywords = {Alleles, Amino Acid Sequence, Animals, Base Sequence, DNA Primers, Drosophila melanogaster, Drosophila Proteins, Frameshift Mutation, Genes, Dominant, Genes, Insect, Molecular Sequence Data, Nuclear Proteins, Phosphoproteins, Point Mutation, Protein Structure, Tertiary, Proteins, RNA Splicing, RNA-Binding Proteins, Sequence Deletion, Sex Determination Analysis}, issn = {0270-7306}, author = {Peng, X and Mount, S M} } @article {49697, title = {Localization of sequences required for size-specific splicing of a small Drosophila intron in vitro.}, journal = {J Mol Biol}, volume = {253}, year = {1995}, month = {1995 Oct 27}, pages = {426-37}, abstract = {

Many introns in Drosophila and other invertebrates are less than 80 nucleotides in length, too small to be recognized by the vertebrate splicing machinery. Comparison of nuclear splicing extracts from human HeLa and Drosophila Kc cells has revealed species-specificity, consistent with the observed size differences. Here we present additional results with the 68 nucleotide fifth intron of the Drosophila myosin heavy chain gene. As observed with the 74 nucleotide second intron of the Drosophila white gene, the wild-type myosin intron is accurately spliced in a homologous extract, and increasing the size by 16 nucleotides both eliminates splicing in the Drosophila extract and allows accurate splicing in the human extract. In contrast to previous results, however, an upstream cryptic 5{\textquoteright} splice site is activated when the wild-type myosin intron is tested in a human HeLa cell nuclear extract, resulting in the removal of a 98 nucleotide intron. The size dependence of splicing in Drosophila extracts is also intron-specific; we noted that a naturally larger (150 nucleotide) intron from the ftz gene is efficiently spliced in Kc cell extracts that do not splice enlarged introns (of 84, 90, 150 or 350 nucleotides) derived from the 74 nucleotide white intron. Here, we have exploited that observation, using a series of hybrid introns to show that a region of 46 nucleotides at the 3{\textquoteright} end of the white intron is sufficient to confer the species-specific size effect. At least two sequence elements within this region, yet distinct from previously described branchpoint and pyrimidine tract signals, are required for efficient splicing of small hybrid introns in vitro.

}, keywords = {Animals, Base Sequence, Cell Line, DNA, Drosophila, Genes, Insect, HeLa Cells, HUMANS, Introns, Molecular Sequence Data, Myosin Heavy Chains, RNA Splicing, Species Specificity}, issn = {0022-2836}, doi = {10.1006/jmbi.1995.0564}, author = {Guo, M and Mount, S M} } @article {38336, title = {Identification of the calcium-binding protein calgranulin in the matrix of struvite stones}, journal = {Journal of endourology / Endourological SocietyJournal of endourology / Endourological Society}, volume = {8}, year = {1994}, note = {http://www.ncbi.nlm.nih.gov/pubmed/8061680?dopt=Abstract}, abstract = {The identification of calcium-binding proteins in urine and kidney stones has led to a closer look at the role of matrix proteins in urolithiasis. We analyzed five struvite stones for protein content and identified two bands (8 and 14 KDa) that were confirmed by gel electrophoresis and amino acid sequencing to be calgranulin. This protein, which is known by several other names, has bacteriostatic antifungal activity. Its role in the formation of struvite stones warrants further investigation.}, keywords = {Amino Acid Sequence, Calcium-Binding Proteins, Cell Adhesion Molecules, Neuronal, Electrophoresis, Enzyme-Linked Immunosorbent Assay, HUMANS, Kidney Calculi, Leukocyte L1 Antigen Complex, Magnesium Compounds, Molecular Sequence Data, Phosphates}, author = {Bennett, J. and Dretler, S. P. and J. Selengut and Orme-Johnson, W. H.} } @article {49699, title = {P element-mediated in vivo deletion analysis of white-apricot: deletions between direct repeats are strongly favored.}, journal = {Genetics}, volume = {136}, year = {1994}, month = {1994 Mar}, pages = {1001-11}, abstract = {

We have isolated and characterized deletions arising within a P transposon, P[hswa], in the presence of P transposase. P[hswa] carries white-apricot (wa) sequences, including a complete copia element, under the control of an hsp70 promoter, and resembles the original wa allele in eye color phenotype. In the presence of P transposase, P[hswa] shows a high overall rate (approximately 3\%) of germline mutations that result in increased eye pigmentation. Of 234 derivatives of P[hswa] with greatly increased eye pigmentation, at least 205 carried deletions within copia. Of these, 201 were precise deletions between the directly repeated 276-nucleotide copia long terminal repeats (LTRs), and four were unique deletions. High rates of transposase-induced precise deletion were observed within another P transposon carrying unrelated 599 nucleotide repeats (yeast 2 mu FLP; recombinase target sites) separated by 5.7 kb. Our observation that P element-mediated deletion formation occurs preferentially between direct repeats suggests general methods for controlling deletion formation.

}, keywords = {Alleles, Animals, Animals, Genetically Modified, Base Sequence, Crosses, Genetic, DNA, DNA Transposable Elements, Drosophila, Eye Color, Female, Genes, Insect, Male, Molecular Sequence Data, Nucleotidyltransferases, PHENOTYPE, Recombination, Genetic, Repetitive Sequences, Nucleic Acid, Sequence Deletion, Transformation, Genetic, Transposases}, issn = {0016-6731}, author = {Kurkulos, M and Weinberg, J M and Roy, D and Mount, S M} } @article {49698, title = {Suppressor U1 snRNAs in Drosophila.}, journal = {Genetics}, volume = {138}, year = {1994}, month = {1994 Oct}, pages = {365-78}, abstract = {

Although the role of U1 small nuclear RNAs (snRNAs) in 5{\textquoteright} splice site recognition is well established, suppressor U1 snRNAs active in intact multicellular animals have been lacking. Here we describe suppression of a 5{\textquoteright} splice site mutation in the Drosophila melanogaster white gene (wDR18) by compensatory changes in U1 snRNA. Mutation of positions -1 and +6 of the 5{\textquoteright} splice site of the second intron (ACG[GTGAGT to ACC]GTGAGC) results in the accumulation of RNA retaining this 74-nucleotide intron in both transfected cells and transgenic flies. U1-3G, a suppressor U1 snRNA which restores base-pairing at position +6 of the mutant intron, increases the ratio of spliced to unspliced wDR18 RNA up to fivefold in transfected Schneider cells and increases eye pigmentation in wDR18 flies. U1-9G, which targets position -1, suppresses wDR18 in transfected cells less well. U1-3G,9G has the same effect as U1-3G although it accumulates to lower levels. Suppression of wDR18 has revealed that the U1b embryonic variant (G134 to U) is active in Schneider cells and pupal eye discs. However, the combination of 9G with 134U leads to reduced accumulation of both U1b-9G and U1b-3G,9G, possibly because nucleotides 9 and 134 both participate in a potential long-range intramolecular base-pairing interaction. High levels of functional U1-3G suppressor reduce both viability and fertility in transformed flies. These results show that, despite the difficulties inherent in stably altering splice site selection in multicellular organisms, it is possible to obtain suppressor U1 snRNAs in flies.

}, keywords = {Alternative Splicing, Animals, Base Sequence, Cell Line, Cell Nucleus, DNA Primers, Drosophila melanogaster, Female, Genes, Suppressor, Genetic Variation, GENOTYPE, Introns, Male, Molecular Sequence Data, Mutagenesis, Site-Directed, Nucleic Acid Conformation, Oligodeoxyribonucleotides, PHENOTYPE, Recombinant Proteins, Ribonucleoprotein, U1 Small Nuclear, RNA, Small Nuclear, Transfection, Transformation, Genetic}, issn = {0016-6731}, author = {Lo, P C and Roy, D and Mount, S M} } @article {49701, title = {Species-specific signals for the splicing of a short Drosophila intron in vitro.}, journal = {Mol Cell Biol}, volume = {13}, year = {1993}, month = {1993 Feb}, pages = {1104-18}, abstract = {

The effects of branchpoint sequence, the pyrimidine stretch, and intron size on the splicing efficiency of the Drosophila white gene second intron were examined in nuclear extracts from Drosophila and human cells. This 74-nucleotide intron is typical of many Drosophila introns in that it lacks a significant pyrimidine stretch and is below the minimum size required for splicing in human nuclear extracts. Alteration of sequences of adjacent to the 3{\textquoteright} splice site to create a pyrimidine stretch was necessary for splicing in human, but not Drosophila, extracts. Increasing the size of this intron with insertions between the 5{\textquoteright} splice site and the branchpoint greatly reduced the efficiency of splicing of introns longer than 79 nucleotides in Drosophila extracts but had an opposite effect in human extracts, in which introns longer than 78 nucleotides were spliced with much greater efficiency. The white-apricot copia insertion is immediately adjacent to the branchpoint normally used in the splicing of this intron, and a copia long terminal repeat insertion prevents splicing in Drosophila, but not human, extracts. However, a consensus branchpoint does not restore the splicing of introns containing the copia long terminal repeat, and alteration of the wild-type branchpoint sequence alone does not eliminate splicing. These results demonstrate species specificity of splicing signals, particularly pyrimidine stretch and size requirements, and raise the possibility that variant mechanisms not found in mammals may operate in the splicing of small introns in Drosophila and possibly other species.

}, keywords = {Animals, Base Sequence, Cell Nucleus, Consensus Sequence, DNA, DNA Transposable Elements, Drosophila, Drosophila Proteins, Electrophoresis, Polyacrylamide Gel, HeLa Cells, HUMANS, Introns, Molecular Sequence Data, Mutation, Peptide Hydrolases, Proteins, Regulatory Sequences, Nucleic Acid, Retroelements, RNA Splicing, Species Specificity}, issn = {0270-7306}, author = {Guo, M and Lo, P C and Mount, S M} } @article {49702, title = {Splicing signals in Drosophila: intron size, information content, and consensus sequences.}, journal = {Nucleic Acids Res}, volume = {20}, year = {1992}, month = {1992 Aug 25}, pages = {4255-62}, abstract = {

A database of 209 Drosophila introns was extracted from Genbank (release number 64.0) and examined by a number of methods in order to characterize features that might serve as signals for messenger RNA splicing. A tight distribution of sizes was observed: while the smallest introns in the database are 51 nucleotides, more than half are less than 80 nucleotides in length, and most of these have lengths in the range of 59-67 nucleotides. Drosophila splice sites found in large and small introns differ in only minor ways from each other and from those found in vertebrate introns. However, larger introns have greater pyrimidine-richness in the region between 11 and 21 nucleotides upstream of 3{\textquoteright} splice sites. The Drosophila branchpoint consensus matrix resembles C T A A T (in which branch formation occurs at the underlined A), and differs from the corresponding mammalian signal in the absence of G at the position immediately preceding the branchpoint. The distribution of occurrences of this sequence suggests a minimum distance between 5{\textquoteright} splice sites and branchpoints of about 38 nucleotides, and a minimum distance between 3{\textquoteright} splice sites and branchpoints of 15 nucleotides. The methods we have used detect no information in exon sequences other than in the few nucleotides immediately adjacent to the splice sites. However, Drosophila resembles many other species in that there is a discontinuity in A + T content between exons and introns, which are A + T rich.

}, keywords = {Animals, Base Sequence, Consensus Sequence, Databases, Factual, Drosophila, Introns, Molecular Sequence Data, RNA Splicing, RNA, Messenger, software}, issn = {0305-1048}, author = {Mount, S M and Burks, C and Hertz, G and Stormo, G D and White, O and Fields, C} } @article {49703, title = {Polyadenylylation in copia requires unusually distant upstream sequences.}, journal = {Proc Natl Acad Sci U S A}, volume = {88}, year = {1991}, month = {1991 Apr 15}, pages = {3038-42}, abstract = {

Retroviruses and related genetic elements generate terminally redundant RNA products by differential polyadenylylation within a long terminal repeat. Expression of the white-apricot (wa) allele of Drosophila melanogaster, which carries an insertion of the 5.1-kilobase retrovirus-like transposable element copia in a small intron, is influenced by signals within copia. By using this indicator, we have isolated a 518-base-pair deletion, 312 base pairs upstream of the copia polyadenylylation site, that is phenotypically like much larger deletions and eliminates RNA species polyadenylylated in copia. This requirement of distant upstream sequences for copia polyadenylylation has implications for the expression of many genetic elements bearing long terminal repeats.

}, keywords = {Animals, Base Sequence, Blotting, Northern, DNA Transposable Elements, Drosophila melanogaster, Eye Color, Molecular Sequence Data, Oligonucleotides, Polymerase Chain Reaction, Regulatory Sequences, Nucleic Acid, Repetitive Sequences, Nucleic Acid, RNA Processing, Post-Transcriptional, RNA, Messenger}, issn = {0027-8424}, author = {Kurkulos, M and Weinberg, J M and Pepling, M E and Mount, S M} } @article {49704, title = {Drosophila melanogaster genes for U1 snRNA variants and their expression during development.}, journal = {Nucleic Acids Res}, volume = {18}, year = {1990}, month = {1990 Dec 11}, pages = {6971-9}, abstract = {

We have cloned and characterized a complete set of seven U1-related sequences from Drosophila melanogaster. These sequences are located at the three cytogenetic loci 21D, 82E, and 95C. Three of these sequences have been previously studied: one U1 gene at 21D which encodes the prototype U1 sequence (U1a), one U1 gene at 82E which encodes a U1 variant with a single nucleotide substitution (U1b), and a pseudogene at 82E. The four previously uncharacterized genes are another U1b gene at 82E, two additional U1a genes at 95C, and a U1 gene at 95C which encodes a new variant (U1c) with a distinct single nucleotide change relative to U1a. Three blocks of 5{\textquoteright} flanking sequence similarity are common to all six full length genes. Using specific primer extension assays, we have observed that the U1b RNA is expressed in Drosophila Kc cells and is associated with snRNP proteins, suggesting that the U1b-containing snRNP particles are able to participate in the process of pre-mRNA splicing. We have also examined the expression throughout Drosophila development of the two U1 variants relative to the prototype sequence. The U1c variant is undetectable by our methods, while the U1b variant exhibits a primarily embryonic pattern reminiscent of the expression of certain U1 variants in sea urchin, Xenopus, and mouse.

}, keywords = {Animals, Base Sequence, Blotting, Southern, Cloning, Molecular, Drosophila melanogaster, Gene Expression Regulation, genes, Genetic Variation, Molecular Sequence Data, Nucleic Acid Conformation, Pseudogenes, Restriction Mapping, RNA, Small Nuclear}, issn = {0305-1048}, author = {Lo, P C and Mount, S M} } @article {49707, title = {Sequence of a cDNA from the Drosophila melanogaster white gene.}, journal = {Nucleic Acids Res}, volume = {18}, year = {1990}, month = {1990 Mar 25}, pages = {1633}, keywords = {Amino Acid Sequence, Animals, Base Sequence, DNA, Drosophila melanogaster, Eye Color, genes, Molecular Sequence Data}, issn = {0305-1048}, author = {Pepling, M and Mount, S M} } @article {49706, title = {Structure and expression of the Drosophila melanogaster gene for the U1 small nuclear ribonucleoprotein particle 70K protein.}, journal = {Mol Cell Biol}, volume = {10}, year = {1990}, month = {1990 Jun}, pages = {2492-502}, abstract = {

A genomic clone encoding the Drosophila U1 small nuclear ribonucleoprotein particle 70K protein was isolated by hybridization with a human U1 small nuclear ribonucleoprotein particle 70K protein cDNA. Southern blot and in situ hybridizations showed that this U1 70K gene is unique in the Drosophila genome, residing at cytological position 27D1,2. Polyadenylated transcripts of 1.9 and 3.1 kilobases were observed. While the 1.9-kilobase mRNA is always more abundant, the ratio of these two transcripts is developmentally regulated. Analysis of cDNA and genomic sequences indicated that these two RNAs encode an identical protein with a predicted molecular weight of 52,879. Comparison of the U1 70K proteins predicted from Drosophila, human, and Xenopus cDNAs revealed 68\% amino acid identity in the most amino-terminal 214 amino acids, which include a sequence motif common to many proteins which bind RNA. The carboxy-terminal half is less well conserved but is highly charged and contains distinctive arginine-rich regions in all three species. These arginine-rich regions contain stretches of arginine-serine dipeptides like those found in transformer, transformer-2, and suppressor-of-white-apricot proteins, all of which have been identified as regulators of mRNA splicing in Drosophila melanogaster.

}, keywords = {Amino Acid Sequence, Animals, Base Sequence, Blotting, Northern, Blotting, Southern, Cloning, Molecular, DNA, Drosophila melanogaster, Gene expression, Gene Library, genes, HUMANS, Molecular Sequence Data, Molecular Weight, Oligonucleotide Probes, Poly A, Ribonucleoproteins, Ribonucleoproteins, Small Nuclear, RNA, RNA, Messenger, Sequence Homology, Nucleic Acid, Xenopus}, issn = {0270-7306}, author = {Mancebo, R and Lo, P C and Mount, S M} } @article {49708, title = {Partial revertants of the transposable element-associated suppressible allele white-apricot in Drosophila melanogaster: structures and responsiveness to genetic modifiers.}, journal = {Genetics}, volume = {118}, year = {1988}, month = {1988 Feb}, pages = {221-34}, abstract = {

The eye color phenotype of white-apricot (wa), a mutant allele of the white locus caused by the insertion of the transposable element copia into a small intron, is suppressed by the extragenic suppressor suppressor-of-white-apricot (su(wa] and enhanced by the extragenic enhancers suppressor-of-forked su(f] and Enhancer-of-white-apricot (E(wa]. Derivatives of wa have been analyzed molecularly and genetically in order to correlate the structure of these derivatives with their response to modifiers. Derivatives in which the copia element is replaced precisely by a solo long terminal repeat (sLTR) were generated in vitro and returned to the germline by P-element mediated transformation; flies carrying this allele within a P transposon show a nearly wild-type phenotype and no response to either su(f) or su(wa). In addition, eleven partial phenotypic revertants of wa were analyzed. Of these, one appears to be a duplication of a large region which includes wa, three are new alleles of su(wa), two are sLTR derivatives whose properties confirm results obtained using transformation, and five are secondary insertions into the copia element within wa. One of these, waR84h, differs from wa by the insertion of the most 3{\textquoteright} 83 nucleotides of the I factor. The five insertion derivatives show a variety of phenotypes and modes of interaction with su[f) and su(wa). The eye pigmentation of waR84h is affected by su(f) and E(wa), but not su(wa). These results demonstrate that copia (as opposed to the interruption of white sequences) is essential for the wa phenotype and its response to genetic modifiers, and that there are multiple mechanisms for the alteration of the wa phenotype by modifiers.

}, keywords = {Alleles, Animals, Base Sequence, DNA Transposable Elements, Drosophila melanogaster, Enhancer Elements, Genetic, GENOTYPE, Molecular Sequence Data, Mutation, Suppression, Genetic}, issn = {0016-6731}, author = {Mount, S M and Green, M M and Rubin, G M} }