To see the other types of publications on this topic, follow the link: Gene expression in C. fusca.

Journal articles on the topic 'Gene expression in C. fusca'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the top 50 journal articles for your research on the topic 'Gene expression in C. fusca.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Browse journal articles on a wide variety of disciplines and organise your bibliography correctly.

1

Deng, Yu, and Stephen S. Fong. "Development and Application of a PCR-Targeted Gene Disruption Method for Studying CelR Function in Thermobifida fusca." Applied and Environmental Microbiology 76, no. 7 (January 22, 2010): 2098–106. http://dx.doi.org/10.1128/aem.02626-09.

Full text
Abstract:
ABSTRACT Thermobifida fusca is a high-G+C-content, thermophilic, Gram-positive soil actinobacterium with high cellulolytic activity. In T. fusca, CelR is thought to act as the primary regulator of cellulase gene expression by binding to a 14-bp inverted repeat [5′-(T)GGGAGCGCTCCC(A)] that is upstream of many known cellulase genes. Previously, the ability to study the roles and regulation of cellulase genes in T. fusca has been limited largely by a lack of established genetic engineering methods for T. fusca. In this study, we developed an efficient procedure for creating precise chromosomal gene disruptions and demonstrated this procedure by generating a celR deletion strain. The celR deletion strain was then characterized using measurements for growth behavior, cellulase activity, and gene expression. The celR deletion strain of T. fusca exhibited a severely crippled growth phenotype with a prolonged lag phase and decreased cell yields for growth on both glucose and cellobiose. While the maximum endoglucanase activity and cellulase activity were not significantly changed, the endoglucanase activity and cellulase activity per cell were highly elevated. Measurements of mRNA transcript levels in both the celR deletion strain and the wild-type strain indicated that the CelR protein potentially acts as a repressor for some genes and as an activator for other genes. Overall, we established and demonstrated a method for manipulating chromosomal DNA in T. fusca that can be used to study the cellulolytic capabilities of this organism. Components of this method may be useful in developing genetic engineering methods for other currently intractable organisms.
APA, Harvard, Vancouver, ISO, and other styles
2

Wei, Yu-Tuo, Qi-Xia Zhu, Zhao-Fei Luo, Fu-Shen Lu, Fa-Zhong Chen, Qing-Yan Wang, Kun Huang, Jian-Zhong Meng, Rong Wang, and Ri-Bo Huang. "Cloning, Expression and Identification of a New Trehalose Synthase Gene from Thermobifida fusca Genome." Acta Biochimica et Biophysica Sinica 36, no. 7 (July 1, 2004): 477–84. http://dx.doi.org/10.1093/abbs/36.7.477.

Full text
Abstract:
Abstract A new open reading frame in Thermobifida fusca sequenced genome was identified to encode a new trehalose synthase, annotated as “glycosidase” in the GenBank database, by bioinformatics searching and experimental validation. The gene had a length of 1830 bp with about 65% GC content and encoded for a new trehalose synthase with 610 amino acids and deduced molecular weight of 66 kD. The high GC content seemed not to affect its good expression in E. coli BL21 in which the target protein could account for as high as 15% of the total cell proteins. The recombinant enzyme showed its optimal activities at 25 °C and pH 6.5 when it converted substrate maltose into trehalose. However it would divert a high proportion of its substrate into glucose when the temperature was increased to 37 °C, or when the enzyme concentration was high Its activity was not inhibited by 5 mM heavy metals such as Cu2+, Mn2+, and Zn2+ but affected by high concentration of glucose. Blasting against the database indicated that amino acid sequence of this protein had maximal 69% homology with the known trehalose synthases, and two highly conserved segments of the protein sequence were identified and their possible linkage with functions was discussed.
APA, Harvard, Vancouver, ISO, and other styles
3

Béki, Emese, István Nagy, Jos Vanderleyden, Szilvia Jäger, László Kiss, László Fülöp, László Hornok, and József Kukolya. "Cloning and Heterologous Expression of a β-d-Mannosidase (EC 3.2.1.25)-Encoding Gene from Thermobifida fusca TM51." Applied and Environmental Microbiology 69, no. 4 (April 2003): 1944–52. http://dx.doi.org/10.1128/aem.69.4.1944-1952.2003.

Full text
Abstract:
ABSTRACT Thermobifida fusca TM51, a thermophilic actinomycete isolated from composted horse manure, was found to produce a number of lignocellulose-degrading hydrolases, including endoglucanases, exoglucanases, endoxylanases, β-xylosidases, endomannanases, and β-mannosidases, when grown on cellulose or hemicellulose as carbon sources. β-Mannosidases (EC 3.2.1.25), although contributing to the hydrolysis of hemicellulose fractions, such as galacto-mannans, constitute a lesser-known group of the lytic enzyme systems due to their low representation in the proteins secreted by hemicellulolytic microorganisms. An expression library of T. fusca, prepared in Streptomyces lividans TK24, was screened for β-mannosidase activity to clone genes coding for mannosidases. One positive clone was identified, and a β-mannosidase-encoding gene (manB) was isolated. Sequence analysis of the deduced amino acid sequence of the putative ManB protein revealed substantial similarity to known mannosidases in family 2 of the glycosyl hydrolase enzymes. The calculated molecular mass of the predicted protein was 94 kDa, with an estimated pI of 4.87. S. lividans was used as heterologous expression host for the putative β-mannosidase gene of T. fusca. The purified gene product obtained from the culture filtrate of S. lividans was then subjected to more-detailed biochemical analysis. Temperature and pH optima of the recombinant enzyme were 53°C and 7.17, respectively. Substrate specificity tests revealed that the enzyme exerts only β-d-mannosidase activity. Its kinetic parameters, determined on para-nitrophenyl β-d-mannopyranoside (pNP-βM) substrate were as follows: Km = 180 μM and V max = 5.96 μmol min−1 mg−1; the inhibition constant for mannose was Ki = 5.5 mM. Glucono-lacton had no effect on the enzyme activity. A moderate trans-glycosidase activity was also observed when the enzyme was incubated in the presence of pNP-αM and pNP-βM; under these conditions mannosyl groups were transferred by the enzyme from pNP-βM to pNP-αM resulting in the synthesis of small amounts (1 to 2%) of disaccharides.
APA, Harvard, Vancouver, ISO, and other styles
4

Morandin, C., A. Hietala, and H. Helanterä. "Vitellogenin and vitellogenin-like gene expression patterns in relation to caste and task in the ant Formica fusca." Insectes Sociaux 66, no. 4 (September 25, 2019): 519–31. http://dx.doi.org/10.1007/s00040-019-00725-9.

Full text
Abstract:
Abstract Social insect colonies are characterized by division of labour, and extensive morphological, physiological and behavioural differences between queens and workers. The storage protein vitellogenin (Vg) affects multiple aspects of social insect life histories, and has been suggested as a key player for caste differentiation and maintenance. Recently, three genes homologous to Vg have been described in the ant Formica exsecta. Their role is currently unclear but their structural variation suggests variable functions. We examined the expression patterns of the conventional Vg and the three Vg-like genes using qRT-PCR in the common black ant Formica fusca between queens and workers, between nurse and foragers workers, and across social contexts (queenless vs. queenright nests), and sampling time. As expected, we found a significant queen caste and nurse task-related increase for the conventional Vg, while Vg-like-C displayed a consistent forager-biased expression pattern. Task (forager vs. nurse) was the only factor that explained expression variation among workers in any of the studied genes. The removal of the queen did not affect expression, although the proportion of fertile nurses increased in queenless nests. The observed expression biases suggest that in Formica fusca, the ancestral duplication has led to alternative social functions for Vg-like genes across castes and tasks. To get a broader picture of the role of gene duplications in social evolution and the roles of Vg-like genes in caste differentiation and maintenance, how these genes achieve these roles at a molecular level need to be investigated further.
APA, Harvard, Vancouver, ISO, and other styles
5

Chen, Wei-Lin, Jo-Chieh Hsu, Chui-Li Lim, Cheng-Yu Chen, and Chao-Hsun Yang. "Expression of the Thermobifida fusca β-1,3-Glucanase in Yarrowia lipolytica and Its Application in Hydrolysis of β-1,3-Glucan from Four Kinds of Polyporaceae." Processes 9, no. 1 (December 29, 2020): 56. http://dx.doi.org/10.3390/pr9010056.

Full text
Abstract:
The gene encoding a thermostable β-1,3-glucanase was cloned from Thermobifida fusca and expressed constitutively by Yarrowia lipolytica using plasmid pYLSC1. The expression level of the recombinant β-1,3-glucanase reached up to 270 U/mL in the culture medium. After a treatment with endo-β-N-acetyl-glucosaminidase H, the recombinant protein appeared as a single protein band, with a molecular size of approximately 66 kDa on the SDS-polyacrylamide gel. The molecular weight was consistent with the size predicted from the nucleotide sequence. The optimum temperature and pH of the transformant β-1,3-glucanase were 60 °C and pH 8.0, respectively. This β-1,3-glucanase was tolerant to 10% methanol, ethanol, and DMSO, retaining 70% activity. The enzyme markedly hydrolyzed Wolfiporia cocos and Pycnoporus sanguineus glucans. The DPPH and ABTS scavenging potential, reducing power and total phenolic contents of these two Polyporaceae hydrolysates, were significantly increased after 18 h of the enzymatic reaction. The present results indicate that T. fusca β-1,3-glucanase from Y. lipolytica transformant (pYLSC1-13g) hydrolyzes W. cocos and P. sanguineus glucans and improves the antioxidant potential of the hydrolysates.
APA, Harvard, Vancouver, ISO, and other styles
6

Vigliotta, Giovanni, Eliana Nutricati, Elisabetta Carata, Salvatore M. Tredici, Mario De Stefano, Paola Pontieri, Domenica Rita Massardo, Maria Vittoria Prati, Luigi De Bellis, and Pietro Alifano. "Clonothrix fusca Roze 1896, a Filamentous, Sheathed, Methanotrophic γ-Proteobacterium." Applied and Environmental Microbiology 73, no. 11 (April 6, 2007): 3556–65. http://dx.doi.org/10.1128/aem.02678-06.

Full text
Abstract:
ABSTRACT Crenothrix polyspora Cohn 1870 and Clonothrix fusca Roze 1896 are two filamentous, sheathed microorganisms exhibiting complex morphological differentiation, whose phylogeny and physiology have been obscure for a long time due to the inability to cultivate them. Very recently, DNA sequencing data from uncultured C. polyspora-enriched material have suggested that Crenothrix is a methane-oxidizing γ-proteobacterium (39). In contrast, the possible ecological function of C. fusca, originally considered a developmental stage of C. polyspora, is unknown. In this study, temporal succession of two filamentous, sheathed microorganisms resembling Cohn's Crenothrix and Roze's Clonothrix was observed by analyzing the microbial community of an artesian well by optical microscopy. Combined culture-based and culture-independent approaches enabled us to assign C. fusca to a novel subgroup of methane-oxidizing γ-proteobacteria distinct from that of C. polyspora. This assignment was supported by (i) methane uptake and assimilation experiments, (ii) ultrastructural data showing the presence in C. fusca cytoplasm of an elaborate membrane system resembling that of methanotrophic γ-proteobacteria, and (iii) sequencing data demonstrating the presence in its genome of a methanol dehydrogenase α subunit-encoding gene (mxaF) and a conventional particulate methane mono-oxygenase α subunit-encoding gene (pmoA) that is different from the unusual pmoA (u-pmoA) of C. polyspora.
APA, Harvard, Vancouver, ISO, and other styles
7

Yang, Chao-Hsun, Kun-I. Lin, Gen-Hung Chen, Yu-Fen Chen, Cheng-Yu Chen, Wei-Lin Chen, and Yu-Chun Huang. "Constitutive Expression of Thermobifida fusca Thermostable Acetylxylan Esterase Gene in Pichia pastoris." International Journal of Molecular Sciences 11, no. 12 (December 15, 2010): 5143–51. http://dx.doi.org/10.3390/ijms11125143.

Full text
APA, Harvard, Vancouver, ISO, and other styles
8

Nayak, Kinshuk C. "Mutational Bias and Gene Expression Level Shape Codon Usage in Thermobifida fusca YX." In Silico Biology 9, no. 5,6 (2009): 337–53. http://dx.doi.org/10.3233/isb-2009-0421.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Ghangas, Gurdev S., and David B. Wilson. "Expression of a Thermomonospora fusca Cellulase Gene in Streptomyces lividans and Bacillus subtilis." Applied and Environmental Microbiology 53, no. 7 (1987): 1470–75. http://dx.doi.org/10.1128/aem.53.7.1470-1475.1987.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Lao, G., and D. B. Wilson. "Cloning, sequencing, and expression of a Thermomonospora fusca protease gene in Streptomyces lividans." Applied and environmental microbiology 62, no. 11 (1996): 4256–59. http://dx.doi.org/10.1128/aem.62.11.4256-4259.1996.

Full text
APA, Harvard, Vancouver, ISO, and other styles
11

Kim, Jeong H., Diana Irwin, and David B. Wilson. "Purification and characterization ofThermobifida fuscaxylanase 10B." Canadian Journal of Microbiology 50, no. 10 (October 1, 2004): 835–43. http://dx.doi.org/10.1139/w04-077.

Full text
Abstract:
Thermobifida fusca grows well on cellulose and xylan, and produces a number of cellulases and xylanases. The gene encoding a previously unstudied endoxylanase, xyl10B, was overexpressed in E. coli, and the protein was purified and characterized. Mature Xyl10B is a 43-kDa glycohydrolase with a short basic domain at the C-terminus. It has moderate thermostability, maintaining 50% of its activity after incubation for 16 h at 62 °C, and is most active between pH 5 and 8. Xyl10B is produced by growth of T. fusca on xylan or Solka Floc but not on pure cellulose. Mass spectroscopic analysis showed that Xyl10B produces xylobiose as the major product from birchwood and oat spelts xylan and that its hydrolysis products differ from those of T. fusca Xyl11A. Xyl10B hydrolyzes various p-nitrophenyl-sugars, including p-nitrophenyl α-D-arabinofuranoside, p-nitrophenyl-β-D-xylobioside, p-nitrophenyl-β-D-xyloside, and p-nitrophenyl-β-D-cellobioside. Xyl11A has higher activity on xylan substrates, but Xyl10B produced more reducing sugars from corn fiber than did Xyl11A.Key words: xylanase, enzyme purification, Thermobifida fusca, family 10 hydrolase.
APA, Harvard, Vancouver, ISO, and other styles
12

Chen, Shaolin, and David B. Wilson. "Proteomic and Transcriptomic Analysis of Extracellular Proteins and mRNA Levels in Thermobifida fusca Grown on Cellobiose and Glucose." Journal of Bacteriology 189, no. 17 (June 29, 2007): 6260–65. http://dx.doi.org/10.1128/jb.00584-07.

Full text
Abstract:
ABSTRACT Thermobifida fusca secretes proteins that carry out plant cell wall degradation. Using two-dimensional electrophoresis, the extracellular proteome of T. fusca grown on cellobiose was compared to that of cells grown on glucose. Extracellular proteins, the expression of which is induced by cellobiose, mainly are cellulases and cellulose-binding proteins. Other major extracellular proteins induced by cellobiose include a xylanase (Xyl10A) and two unknown proteins, the C-terminal regions of which are homologous to a lytic transglycosylase goose egg white lysozyme domain and an NLPC_P60 domain (which defines a family of cell wall peptidases), respectively. Transcriptional analysis of genes encoding cellobiose-induced proteins suggests that their expression is controlled at the transcriptional level and that their expression also is induced by cellulose. Some other major extracellular proteins produced by T. fusca grown on both cellobiose and glucose include Lam81A and three unknown proteins that are homologous to aminopeptidases and xylanases or that contain a putative NLPC_P60 domain.
APA, Harvard, Vancouver, ISO, and other styles
13

Jung, E. D., G. Lao, D. Irwin, B. K. Barr, A. Benjamin, and D. B. Wilson. "DNA sequences and expression in Streptomyces lividans of an exoglucanase gene and an endoglucanase gene from Thermomonospora fusca." Applied and Environmental Microbiology 59, no. 9 (1993): 3032–43. http://dx.doi.org/10.1128/aem.59.9.3032-3043.1993.

Full text
APA, Harvard, Vancouver, ISO, and other styles
14

Park, Sooyeon, Yong-Taek Jung, Chul-Hyung Kang, Ja-Min Park, and Jung-Hoon Yoon. "Thalassotalea ponticola sp. nov., isolated from seawater, reclassification of Thalassomonas fusca as Thalassotalea fusca comb. nov. and emended description of the genus Thalassotalea." International Journal of Systematic and Evolutionary Microbiology 64, Pt_11 (November 1, 2014): 3676–82. http://dx.doi.org/10.1099/ijs.0.067611-0.

Full text
Abstract:
A Gram-stain-negative, aerobic, non-flagellated and rod-shaped or ovoid bacterial strain, designated GJSW-36T, was isolated from seawater at Geoje island in the South Sea, South Korea. Strain GJSW-36T grew optimally at pH 7.0–8.0, at 25 °C and in the presence of 2 % (w/v) NaCl. A neighbour-joining phylogenetic tree based on 16S rRNA gene sequences showed that strain GJSW-36T fell within the clade comprising the type strains of species of the genus Thalassotalea and Thalassomonas fusca . Strain GJSW-36T exhibited 16S rRNA gene sequence similarity values of 94.2–96.0 % to the type strains of species of the genus Thalassotalea and Thalassomonas fusca and of 93.8–94.5 % to the type strains of the other species of the genus Thalassomonas . Strain GJSW-36T contained ubiquinone-8 (Q-8) as the predominant ubiquinone and summed feature 3 (C16 : 1ω7c and/or C16 : 1ω6c), C17 : 1ω8c and C16 : 0 as the major fatty acids. The major polar lipids of strain GJSW-36T were phosphatidylglycerol and phosphatidylethanolamine. The DNA G+C content of strain GJSW-36T was 45.1 mol%. Differential phenotypic properties, together with the phylogenetic distinctiveness, demonstrated that strain GJSW-36T is separated from species of the genus Thalassotalea and Thalassomonas fusca . On the basis of the data presented, strain GJSW-36T is considered to represent a novel species of the genus Thalassotalea , for which the name Thalassotalea ponticola sp. nov. is proposed. The type strain is GJSW-36T ( = KCTC 42155T = CECT 8656T). From this study, it is also proposed that Thalassomonas fusca should be reclassified as a member of the genus Thalassotalea and the description of the genus Thalassotalea is emended.
APA, Harvard, Vancouver, ISO, and other styles
15

Cheepudom, Lin, Lee, and Meng. "Characterization of a Novel Thermobifida fusca Bacteriophage P318." Viruses 11, no. 11 (November 8, 2019): 1042. http://dx.doi.org/10.3390/v11111042.

Full text
Abstract:
Thermobifida fusca is of biotechnological interest due to its ability to produce an array of plant cell wall hydrolytic enzymes. Nonetheless, only one T. fusca bacteriophage with genome information has been reported to date. This study was aimed at discovering more relevant bacteriophages to expand the existing knowledge of phage diversity for this host species. With this end in view, a thermostable T. fusca bacteriophage P318, which belongs to the Siphoviridae family, was isolated and characterized. P318 has a double-stranded DNA genome of 48,045 base pairs with 3′-extended COS ends, on which 52 putative ORFs are organized into clusters responsible for the order of genome replication, virion morphogenesis, and the regulation of the lytic/lysogenic cycle. In comparison with T. fusca and the previously discovered bacteriophage P1312, P318 has a much lower G+C content in its genome except at the region encompassing ORF42, which produced a protein with unknown function. P1312 and P318 share very few similarities in their genomes except for the regions encompassing ORF42 of P318 and ORF51 of P1312 that are homologous. Thus, acquisition of ORF42 by lateral gene transfer might be an important step in the evolution of P318.
APA, Harvard, Vancouver, ISO, and other styles
16

Vaz-Moreira, Ivone, Cátia Faria, Ana R. Lopes, Liselott A. Svensson, Edward R. B. Moore, Olga C. Nunes, and Célia M. Manaia. "Shinella fusca sp. nov., isolated from domestic waste compost." International Journal of Systematic and Evolutionary Microbiology 60, no. 1 (January 1, 2010): 144–48. http://dx.doi.org/10.1099/ijs.0.009498-0.

Full text
Abstract:
A bacterium, designated strain DC-196T, isolated from kitchen refuse compost was analysed by using a polyphasic approach. Strain DC-196T was characterized as a Gram-negative short rod that was catalase- and oxidase-positive, and able to grow at 10–40 °C, pH 6–9 and in NaCl concentrations as high as 3 %. Chemotaxonomically, C18 : 1 was observed to be the predominant cellular fatty acid and ubiquinone 10 (Q10) was the predominant respiratory quinone. The G+C content of the genomic DNA was determined to be 66 mol%. On the basis of the genotypic, phenotypic and chemotaxonomic characteristics, strain DC-196T was assigned to the genus Shinella, although with distinctive features. At the time of writing, 16S rRNA gene sequence similarities of 97.6–96.8 % and the low DNA–DNA hybridization values of 38.2–32.2 % with the type strains of the three recognized Shinella species confirmed that strain DC-196T represents a novel species of the genus, for which the name Shinella fusca sp. nov. is proposed (type strain DC-196T=CCUG 55808T=LMG 24714T).
APA, Harvard, Vancouver, ISO, and other styles
17

Ghangas, G. S., Y. J. Hu, and D. B. Wilson. "Cloning of a Thermomonospora fusca xylanase gene and its expression in Escherichia coli and Streptomyces lividans." Journal of Bacteriology 171, no. 6 (1989): 2963–69. http://dx.doi.org/10.1128/jb.171.6.2963-2969.1989.

Full text
APA, Harvard, Vancouver, ISO, and other styles
18

Rodriguez Buitrago, Jhon Alexander, Thomas Klünemann, Wulf Blankenfeldt, and Anett Schallmey. "Expression, purification and crystal structure determination of a ferredoxin reductase from the actinobacterium Thermobifida fusca." Acta Crystallographica Section F Structural Biology Communications 76, no. 8 (July 28, 2020): 334–40. http://dx.doi.org/10.1107/s2053230x2000922x.

Full text
Abstract:
The ferredoxin reductase FdR9 from Thermobifida fusca, a member of the oxygenase-coupled NADH-dependent ferredoxin reductase (FNR) family, catalyses electron transfer from NADH to its physiological electron acceptor ferredoxin. It forms part of a putative three-component cytochrome P450 monooxygenase system in T. fusca comprising CYP222A1 and the [3Fe–4S]-cluster ferredoxin Fdx8 as well as FdR9. Here, FdR9 was overexpressed and purified and its crystal structure was determined at 1.9 Å resolution. The overall structure of FdR9 is similar to those of other members of the FNR family and is composed of an FAD-binding domain, an NAD-binding domain and a C-terminal domain. Activity measurements with FdR9 confirmed a strong preference for NADH as the cofactor. Comparison of the FAD- and NAD-binding domains of FdR9 with those of other ferredoxin reductases revealed the presence of conserved sequence motifs in the FAD-binding domain as well as several highly conserved residues involved in FAD and NAD cofactor binding. Moreover, the NAD-binding site of FdR9 contains a modified Rossmann-fold motif, GxSxxS, instead of the classical GxGxxG motif.
APA, Harvard, Vancouver, ISO, and other styles
19

Ventura, Carlo, and Margherita Maioli. "Protein Kinase C Control of Gene Expression." Critical Reviews™ in Eukaryotic Gene Expression 11, no. 1-3 (2001): 26. http://dx.doi.org/10.1615/critreveukargeneexpr.v11.i1-3.120.

Full text
APA, Harvard, Vancouver, ISO, and other styles
20

Asselah, T., I. Bieche, A. Sabbagh, P. Bedossa, R. Moreau, D. Valla, M. Vidaud, and P. Marcellin. "Gene expression and hepatitis C virus infection." Gut 58, no. 6 (December 11, 2008): 846–58. http://dx.doi.org/10.1136/gut.2008.166348.

Full text
APA, Harvard, Vancouver, ISO, and other styles
21

Feng, Dingxia, Jenifer Saldanha, Qi Ye, and Jo Anne Powell-Coffman. "Oxygen-sensitive gene expression in C. elegans." Developmental Biology 356, no. 1 (August 2011): 257–58. http://dx.doi.org/10.1016/j.ydbio.2011.05.488.

Full text
APA, Harvard, Vancouver, ISO, and other styles
22

Xie, Yue, Changmin Chen, and David A. Hume. "Transcriptional Regulation of c-fms Gene Expression." Cell Biochemistry and Biophysics 34, no. 1 (2001): 001–16. http://dx.doi.org/10.1385/cbb:34:1:001.

Full text
APA, Harvard, Vancouver, ISO, and other styles
23

Su, Lingqia, Ruoyu Hong, and Jing Wu. "Enhanced extracellular expression of gene-optimized Thermobifida fusca cutinase in Escherichia coli by optimization of induction strategy." Process Biochemistry 50, no. 7 (July 2015): 1039–46. http://dx.doi.org/10.1016/j.procbio.2015.03.023.

Full text
APA, Harvard, Vancouver, ISO, and other styles
24

Deng, Yu, and Xiaojuan Zhang. "DtxR, an iron-dependent transcriptional repressor that regulates the expression of siderophore gene clusters in Thermobifida fusca." FEMS Microbiology Letters 362, no. 3 (February 1, 2015): 1–6. http://dx.doi.org/10.1093/femsle/fnu053.

Full text
APA, Harvard, Vancouver, ISO, and other styles
25

Sarasin-Filipowicz, Magdalena, and Markus H. Heim. "Interferon-induced gene expression in chronic hepatitis C." Future Virology 5, no. 1 (January 2010): 25–31. http://dx.doi.org/10.2217/fvl.09.67.

Full text
APA, Harvard, Vancouver, ISO, and other styles
26

Kumamoto, Carol A. "Niche-specific gene expression during C. albicans infection." Current Opinion in Microbiology 11, no. 4 (August 2008): 325–30. http://dx.doi.org/10.1016/j.mib.2008.05.008.

Full text
APA, Harvard, Vancouver, ISO, and other styles
27

Mukai, Harumi Y., Tomoko Kono, Hozumi Motohashi, Masayuki Yamamoto, and Hiroshi Kojima. "c-Myb Regulates CD9 Gene Expression during Megakaryopoiesis." Blood 110, no. 11 (November 16, 2007): 1245. http://dx.doi.org/10.1182/blood.v110.11.1245.1245.

Full text
Abstract:
Abstract The nuclear proto-oncogene c-myb plays crucial roles in the growth, survival, and differentiation of hematopoietic cells. We previously generated through insertion mutagenesis a c-myb gene knockdown (KD) line of mice. In the mice transgene was inserted 77-kb upstream of the c-myb gene and c-Myb expression was markedly decreased in megakaryocyte-erythrocyte lineage-restricted progenitors (MEPs) of the homozygous knockdown mutant mice (c-myb KD mice). The c-myb KD mice exhibited anemia, thrombocythemia, and splenomegaly and these abnormalities were reproducible in a co-culture assay of MEPs with OP9 cells, but abrogated by the retroviral expression of c-Myb in MEPs. To understand the transcriptional program that accompanies the decline of c-myb gene expression, we performed DNA microarray analysis with MEPs and identified 74 genes that are upregulated and 36 genes that are downregulated in the c-myb KD mice. Of these genes, expression levels 15 genes are actually changed significantly in bone marrow cells of the c-myb KD mice. These genes harbor c-Myb recognition elements in their regulatory regions. Especially, we found that the CD9 expression was upregulated in the c-myb KD mice. Reverse correlation of c-Myb expression with the CD9 gene expression was verified using a luciferase reporter assay and chromatin immunoprecipitation assay. Agonistic antibody of CD9 stimulated megakaryocytic colony formation. On the contrary, upon the bone marrow suppression with 5-fluorouracil recovery of platelet number was delayed in the CD9-null mice. Furthermore, proplatelet formation was impaired when we used CD9-null mouse megakaryocytes, and the size of proplatelets was smaller than those generated by wild-type megakaryocytes. These results thus demonstrate that c-Myb suppresses the CD9 expression in a steady-state condition, while in the stress megakaryopoiesis CD9 is derepressed and acts to induce the megakaryopoiesis. Elucidation of c-myb-based transcription network seems to be of important to understand the megakaryocytic differentiation.
APA, Harvard, Vancouver, ISO, and other styles
28

Yong-Xia, Shi, Lv Lei, Xu Wei, Yuan Mei-Jin, and Pang Yi. "Expression ofvip2A(c) gene fromBacillus thuringiensisin insect cells." Chinese Journal of Agricultural Biotechnology 3, no. 3 (December 2006): 223–29. http://dx.doi.org/10.1079/cjb2006114.

Full text
Abstract:
AbstractRecombinant Bac-GV2 DNA was obtained by inserting a fusedgfpgene with theBacillus thuringiensis vip2A(c) gene encoding a possible enzymatic component under the control of the polyhedrin gene promoter of the baculovirusAutographa californica multicapsid nucleopolyhedrovirus(AcMNPV). TheTrichoplusia nicell line TnHi5 was transfected with Bac-GFP and Bac-GV2 DNAs respectively. Fluorescent cells expressing the fusion protein GV2 were much fewer than those expressing green fluorescent protein (GFP) alone, and did not obviously increase in number from 2 to 5 days after transfection. This result showed that the Vip2A fusion protein might have an ADP-ribosylating activity on cell skeleton actin, exerting an influence on the production and diffusion of the budded virus from insect cells.
APA, Harvard, Vancouver, ISO, and other styles
29

Liebetrau, W., A. Budde, A. Savoia, F. Grummt, and H. Hoehn. "P53 Activates Fanconi Anemia Group C Gene Expression." Human Molecular Genetics 6, no. 2 (February 1, 1997): 277–83. http://dx.doi.org/10.1093/hmg/6.2.277.

Full text
APA, Harvard, Vancouver, ISO, and other styles
30

SKRZYPCZAK, MACIEJ, JULITA ZIELEWICZ, STANISŁAW WINIARCZYK, JACEK WOJCIEROWSKI, ANNA KWAŚNIEWSKA, and JERZY JAKOWICKI. "Expression of c-MYC gene in neoplastic endometrium." PRZEGLĄD GINEKOLOGICZNO-POŁOŻNICZY 5, no. 1-1 (February 23, 2005): 15–20. http://dx.doi.org/10.1066/s10014040059.

Full text
APA, Harvard, Vancouver, ISO, and other styles
31

Hill, A. A. "Genomic Analysis of Gene Expression in C. elegans." Science 290, no. 5492 (October 27, 2000): 809–12. http://dx.doi.org/10.1126/science.290.5492.809.

Full text
APA, Harvard, Vancouver, ISO, and other styles
32

Su, Tsung-Sheng, Ling-Huang Lin, Wing-Yiu Lui, Chungming Chang, Chen-Kung Chou, Ling-Pei Ting, Cheng-Po Hu, Shou-Hwa Han, and Fang-Ku P'eng. "Expression of c-myc gene in human hepatoma." Biochemical and Biophysical Research Communications 132, no. 1 (October 1985): 264–68. http://dx.doi.org/10.1016/0006-291x(85)91017-4.

Full text
APA, Harvard, Vancouver, ISO, and other styles
33

Woloschak, M., J. L. Roberts, and K. Post. "c-myc, c-fos, and c-myb gene expression in human pituitary adenomas." Journal of Clinical Endocrinology & Metabolism 79, no. 1 (July 1994): 253–57. http://dx.doi.org/10.1210/jcem.79.1.8027238.

Full text
APA, Harvard, Vancouver, ISO, and other styles
34

Woloschak, M. "c-myc, c-fos, and c-myb gene expression in human pituitary adenomas." Journal of Clinical Endocrinology & Metabolism 79, no. 1 (July 1, 1994): 253–57. http://dx.doi.org/10.1210/jc.79.1.253.

Full text
APA, Harvard, Vancouver, ISO, and other styles
35

Martin, Stephen J., Heikki Helanterä, and Falko P. Drijfhout. "Is parasite pressure a driver of chemical cue diversity in ants?" Proceedings of the Royal Society B: Biological Sciences 278, no. 1705 (July 7, 2010): 496–503. http://dx.doi.org/10.1098/rspb.2010.1047.

Full text
Abstract:
Parasites and pathogens are possibly key evolutionary forces driving recognition systems. However, empirical evidence remains sparse. The ubiquitous pioneering ant Formica fusca is exploited by numerous socially parasitic ant species. We compared the chemical cue diversity, egg and nest mate recognition abilities in two Finnish and two UK populations where parasite pressure is high or absent, respectively. Finnish populations had excellent egg and nest mate discrimination abilities, which were lost in the UK populations. The loss of discrimination behaviour correlates with a loss in key recognition compounds (C 25 -dimethylalkanes). This was not owing to genetic drift or different ecotypes since neutral gene diversity was the same in both countries. Furthermore, it is known that the cuticular hydrocarbon profiles of non-host ant species remain stable between Finland and the UK. The most parsimonious explanation for the striking difference in the cue diversity (number of C 25 -dimethylalkanes isomers) between the UK and Finland populations is the large differences in parasite pressure experienced by F. fusca in the two countries. These results have strong parallels with bird (cuckoo) studies and support the hypothesis that parasites are driving recognition cue diversity.
APA, Harvard, Vancouver, ISO, and other styles
36

Park, Yang-Nim, and Joachim Morschhäuser. "Tetracycline-Inducible Gene Expression and Gene Deletion in Candida albicans." Eukaryotic Cell 4, no. 8 (August 2005): 1328–42. http://dx.doi.org/10.1128/ec.4.8.1328-1342.2005.

Full text
Abstract:
ABSTRACT The genetic analysis of Candida albicans, the major fungal pathogen of humans, is hampered by its diploid genome, the absence of a normal sexual cycle, and a nonstandard codon usage. Although effective methods to study gene function have been developed in the past years, systems to control gene expression in C. albicans are limited. We have established a system that allows induction of gene expression in C. albicans by the addition of tetracycline (Tet). By fusing genetically modified versions of the reverse Tet repressor from Escherichia coli and the transcription activation domain of the Gal4 protein from Saccharomyces cerevisiae, a C. albicans-adapted reverse Tet-dependent transactivator (rtTA) was created that was expressed from the constitutive ADH1 or the opaque-specific OP4 promoter. To monitor Tet-inducible gene expression, the caGFP reporter gene was placed under the control of a Tet-dependent promoter, obtained by fusing a minimal promoter from C. albicans to seven copies of the Tet operator sequence. Fluorescence of the cells demonstrated that gene expression could be efficiently induced by the addition of doxycycline in yeast, hyphal, and opaque cells of C. albicans. The Tet-inducible gene expression system was then used to manipulate the behavior of the various growth forms of C. albicans. Tet-induced expression of a dominant-negative CDC42 allele resulted in growth arrest as large, multinucleate cells. Filamentous growth was efficiently inhibited under all tested hyphal-growth-promoting conditions by Tet-inducible expression of the NRG1 repressor. Tet-induced expression of the MTL a 1 gene in opaque cells of an MTLα strain forced the cells to switch to the white phase, whereas Tet-induced expression of the MTL a 2 transcription factor induced shmooing. When the ecaFLP gene, encoding the site-specific recombinase FLP, was placed under the control of the Tet-dependent promoter, Tet-inducible deletion of genes which were flanked by the FLP target sequences was achieved with high efficiency to generate conditional null mutants. In combination with the dominant selection marker caSAT1, the Tet-inducible gene expression system was also applied in C. albicans wild-type strains, including drug-resistant clinical isolates that overexpressed the MDR1, CDR1, and CDR2 multidrug efflux pumps. This system, therefore, allows a growth medium-independent, Tet-inducible expression and deletion of genes in C. albicans and provides a convenient, versatile new tool to study gene function and manipulate cellular behavior in this model pathogenic fungus.
APA, Harvard, Vancouver, ISO, and other styles
37

Deng, Yi-Bin, Le-Gen Nong, Zuo-Ren Liang, Liang Zhang, Yu-Hua Qin, and Ping He. "Hepatitis C virus gene-specific locked nucleic acid enzyme significantly inhibits C gene expression in vitro." World Chinese Journal of Digestology 22, no. 14 (2014): 1992. http://dx.doi.org/10.11569/wcjd.v22.i14.1992.

Full text
APA, Harvard, Vancouver, ISO, and other styles
38

Ding, Hao, Xiaoli Wu, and Andras Nagy. "Mice with Cre recombinase activatable PDGF-C expression." genesis 32, no. 2 (February 2002): 181–83. http://dx.doi.org/10.1002/gene.10065.

Full text
APA, Harvard, Vancouver, ISO, and other styles
39

Zahn, Laura M. "Gene expression can point to disease risk." Science 364, no. 6445 (June 13, 2019): 1044.3–1045. http://dx.doi.org/10.1126/science.364.6445.1044-c.

Full text
APA, Harvard, Vancouver, ISO, and other styles
40

Choi, H. S. "Induction of c-fos and c-jun gene expression by phenolic antioxidants." Molecular Endocrinology 7, no. 12 (December 1, 1993): 1596–602. http://dx.doi.org/10.1210/me.7.12.1596.

Full text
APA, Harvard, Vancouver, ISO, and other styles
41

GARZINO-DEMO, ALFREDO, SURESH K. ARYA, ANTHONY L. DEVICO, FIORENZA COCCHI, PAOLO LUSSO, and ROBERT C. GALLO. "C-C Chemokine RANTES and HIV Long Terminal Repeat-Driven Gene Expression." AIDS Research and Human Retroviruses 13, no. 16 (November 1997): 1367–71. http://dx.doi.org/10.1089/aid.1997.13.1367.

Full text
APA, Harvard, Vancouver, ISO, and other styles
42

Choi, H. S., and D. D. Moore. "Induction of c-fos and c-jun gene expression by phenolic antioxidants." Molecular Endocrinology 7, no. 12 (December 1993): 1596–602. http://dx.doi.org/10.1210/mend.7.12.8145765.

Full text
APA, Harvard, Vancouver, ISO, and other styles
43

Bynum, James A., Tiffani Chance, Michael Adam Meledeo, and Heather F. Pidcoke. "Gene Expression Profiling Reveals Key Mitochondrial Gene Changes in Stored Platelets." Blood 126, no. 23 (December 3, 2015): 3559. http://dx.doi.org/10.1182/blood.v126.23.3559.3559.

Full text
Abstract:
Abstract Introduction As of June 2015 the FDA has approved an alternative procedure under 21 CFR 640.120 that allows for storage of apheresis platelets at refrigerator temperature (1-6 C; 4°C) without agitation for up to 3 days for use in the resuscitation of actively bleeding patients. Understanding underlying mechanisms responsible for enhanced hemostatic function at 4°C will be critical for such improvements in platelet transfusion. We hypothesized that 4°C platelets display better mitochondrial respiratory function for up to 7 days compared to standard 5-day RT platelets and that mitochondrial gene expression differences between RT and 4°C -stored platelets will correlate with mitochondrial function. Methods Platelets were collected from healthy donors by apheresis according to an IRB-approved protocol. Apheresis platelets (AP) were rested for 1 h before allocation into platelet minibags (Blood Cell Storage, Seattle, WA) and stored for 4 storage durations (Baseline (BL), Day 3, 5, and 7). Mitochondrial respiration, maximal oxygen utilization, and individual mitochondrial complex-dependent respiration were assessed with high-resolution respirometry (O2k, Oroboros). Mitochondrial ROS generation in response to storage condition or stimulation (to assess oxidative burst capacity as a measure of function) was visualized with fluorescent imaging and assayed with flow cytometry using a superoxide stain (Life Technologies). Total RNA was extracted both immediately following apheresis (BL) and on Day 5 from RT and 4°C-stored platelets using Trizol (Molecular Research Center, Cincinnati, OH) after centrifuging the platelets at 900 x g for 10 min. Platelet RNA was quantified using the NanoDrop 2000. RNA quality was examined using gel electrophoresis with the Reliant Gel System (Cambrex, Rockland, ME). Platelet mitochondrial gene expression analysis was evaluated using the 96-well RT2 Profiler PCR Array (Qiagen, Valencia, CA) which profiled 84 mitochondria-focused targets and 12 control genes per sample. Gene expression data analysis was based on the ΔΔCt method with normalization of the raw data to housekeeping genes located on each 96-well plate. Results Mitochondrial respiration was lower in platelets stored at 4°C compared to RT on Days 3, 5, and 7 (Day 5= -57%±0.3; P < 0.05), demonstrating that refrigeration slows metabolism. Additionally, maximal mitochondrial oxygen utilization (electron transport system capacity) was better preserved in platelets stored at 4°C (Figure 1). Fluorescent imaging and flow cytometry demonstrated that mROS generation was higher in RT-stored platelets compared to 4°C, reflecting mitochondrial damage. Mitochondrial burst during de novo mROS generation due to stimulation was also preserved at 4°C. Mitochondrial gene expression studies revealed distinct differences in expression profiles for 4°C versus RT-stored platelets after 5 days of storage when normalized to BL measures. Storage at 4°C resulted in significantly greater preservation of 15 gene products at Day 5 (P<0.05). In contrast, Day 5 RT samples resulted in a marked decrease or loss of gene products when compared to BL levels of gene expression (P<0.05). Discussion Platelet mitochondrial respiratory function (mitochondrial respiration and maximal oxygen utilization) decreased in RT-stored platelets over 7 days, but the impairment was attenuated by 4°C storage. We previously noted that intracellular ROS flux was higher at room temperature, and here the gene expression analysis in combination with oximetry data showed that mitochondrial damage is likely responsible. Furthermore, gene expression profiling of mitochondrial-related genes revealed that distinct differences exist in key mitochondrial genes between the storage conditions. This work illustrates that 4°C storage of platelets preserved and enhanced critical mitochondrial genes compared to RT; this finding combined with improved mitochondrial respiratory measures and reduced ROS demonstrates a significant improvement in current efforts to mitigate platelet dysfunction. Figure 1. Maximal respiration induced by titration of the protonophore FCCP (carbonyl cyanide p-(trifluoromethoxy) phenylhydrazone) demonstrated a significant decrease in mitochondrial capacity (indicating loss of function) in RT-stored samples compared to 4 °C-stored samples by Day 7. Values are mean ± SD (n=7); *P<0.001 compared to RT. Figure 1. Maximal respiration induced by titration of the protonophore FCCP (carbonyl cyanide p-(trifluoromethoxy) phenylhydrazone) demonstrated a significant decrease in mitochondrial capacity (indicating loss of function) in RT-stored samples compared to 4 °C-stored samples by Day 7. Values are mean ± SD (n=7); *P<0.001 compared to RT. Disclosures No relevant conflicts of interest to declare.
APA, Harvard, Vancouver, ISO, and other styles
44

Palkó, Enikő, Szilárd Póliska, Zsuzsanna Csákányi, Gábor Katona, Tamás Karosi, Frigyes Helfferich, András Penyige, and István Sziklai. "The c-MYC Protooncogene Expression in Cholesteatoma." BioMed Research International 2014 (2014): 1–6. http://dx.doi.org/10.1155/2014/639896.

Full text
Abstract:
Cholesteatoma is an epidermoid cyst, which is most frequently found in the middle ear. The matrix of cholesteatoma is histologically similar to the matrix of the epidermoid cyst of the skin (atheroma); their epithelium is characterized by hyperproliferation. The c-MYC protooncogene located on chromosome 8q24 encodes a transcription factor involved in the regulation of cell proliferation and differentiation. Previous studies have found aneuploidy of chromosome 8, copy number variation of c-MYC gene, and the presence of elevated level c-MYC protein in cholesteatoma. In this study we have compared the expression of c-MYC gene in samples taken from the matrix of 26 acquired cholesteatomas (15 children and 11 adults), 15 epidermoid cysts of the skin (atheromas; head and neck region) and 5 normal skin samples (retroauricular region) using RT-qPCR, providing the first precise measurement of the expression of c-MYC gene in cholesteatoma. We have found significantly elevated c-MYC gene expression in cholesteatoma compared to atheroma and to normal skin samples. There was no significant difference, however, in c-MYC gene expression between cholesteatoma samples of children and adults. The significant difference in c-MYC gene expression level in cholesteatoma compared to that of atheroma implies a more prominent hyperproliferative phenotype which may explain the clinical behavior typical of cholesteatoma.
APA, Harvard, Vancouver, ISO, and other styles
45

LU, HUADING, GANG HOU, YONGKAI ZHANG, YUHU DAI, and HUIQING ZHAO. "c-Jun transactivates Puma gene expression to promote osteoarthritis." Molecular Medicine Reports 9, no. 5 (February 24, 2014): 1606–12. http://dx.doi.org/10.3892/mmr.2014.1981.

Full text
APA, Harvard, Vancouver, ISO, and other styles
46

Puckelwartz, Megan J., Frederic F. S. Depreux, and Elizabeth M. McNally. "Gene expression, chromosome position and lamin A/C mutations." Nucleus 2, no. 3 (May 2011): 162–67. http://dx.doi.org/10.4161/nucl.2.3.16003.

Full text
APA, Harvard, Vancouver, ISO, and other styles
47

Sharma, Shilpa, and Gurudutta Gangenahalli. "Gene Expression Profiling of Human c-Kit Mutant D816V." Journal of Cancer Therapy 07, no. 06 (2016): 439–54. http://dx.doi.org/10.4236/jct.2016.76046.

Full text
APA, Harvard, Vancouver, ISO, and other styles
48

Sugi, Takuma, Yukuo Nishida, and Ikue Mori. "Gene expression dynamics that regulates C. elegans behavioral memory." Neuroscience Research 71 (September 2011): e46. http://dx.doi.org/10.1016/j.neures.2011.07.196.

Full text
APA, Harvard, Vancouver, ISO, and other styles
49

Vogel, R. F., W. Gaier, and W. P. Hammes. "Expression of the lipase gene fromStaphylococcus hyicusinLactobacillus curvantsLc2-c." FEMS Microbiology Letters 69, no. 3 (June 1990): 289–92. http://dx.doi.org/10.1111/j.1574-6968.1990.tb04246.x.

Full text
APA, Harvard, Vancouver, ISO, and other styles
50

Hendriks, Gert-Jan, Dimos Gaidatzis, Florian Aeschimann, and Helge Großhans. "Extensive Oscillatory Gene Expression during C. elegans Larval Development." Molecular Cell 53, no. 3 (February 2014): 380–92. http://dx.doi.org/10.1016/j.molcel.2013.12.013.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography