首页 | 本学科首页   官方微博 | 高级检索  
文章检索
  按 检索   检索词:      
出版年份:   被引次数:   他引次数: 提示:输入*表示无穷大
  收费全文   1980篇
  免费   155篇
  2135篇
  2023年   8篇
  2022年   13篇
  2021年   34篇
  2020年   17篇
  2019年   23篇
  2018年   40篇
  2017年   27篇
  2016年   63篇
  2015年   83篇
  2014年   99篇
  2013年   114篇
  2012年   155篇
  2011年   129篇
  2010年   86篇
  2009年   82篇
  2008年   120篇
  2007年   121篇
  2006年   127篇
  2005年   98篇
  2004年   99篇
  2003年   110篇
  2002年   88篇
  2001年   11篇
  2000年   17篇
  1999年   30篇
  1998年   26篇
  1997年   25篇
  1996年   20篇
  1995年   14篇
  1994年   17篇
  1993年   10篇
  1992年   9篇
  1991年   10篇
  1990年   11篇
  1989年   6篇
  1988年   13篇
  1987年   13篇
  1986年   11篇
  1985年   12篇
  1984年   13篇
  1983年   12篇
  1982年   14篇
  1981年   14篇
  1980年   15篇
  1978年   6篇
  1977年   6篇
  1976年   8篇
  1975年   10篇
  1974年   6篇
  1973年   6篇
排序方式: 共有2135条查询结果,搜索用时 15 毫秒
101.
The emergence of multicellularity is strongly correlated with the expansion of tyrosine kinases, a conserved family of signaling enzymes that regulates pathways essential for cell-to-cell communication. Although tyrosine kinases have been classified from several model organisms, a molecular-level understanding of tyrosine kinase evolution across all holozoans is currently lacking. Using a hierarchical sequence constraint-based classification of diverse holozoan tyrosine kinases, we construct a new phylogenetic tree that identifies two ancient clades of cytoplasmic and receptor tyrosine kinases separated by the presence of an extended insert segment in the kinase domain connecting the D and E-helices. Present in nearly all receptor tyrosine kinases, this fast-evolving insertion imparts diverse functionalities, such as post-translational modification sites and regulatory interactions. Eph and EGFR receptor tyrosine kinases are two exceptions which lack this insert, each forming an independent lineage characterized by unique functional features. We also identify common constraints shared across multiple tyrosine kinase families which warrant the designation of three new subgroups: Src module (SrcM), insulin receptor kinase-like (IRKL), and fibroblast, platelet-derived, vascular, and growth factor receptors (FPVR). Subgroup-specific constraints reflect shared autoinhibitory interactions involved in kinase conformational regulation. Conservation analyses describe how diverse tyrosine kinase signaling functions arose through the addition of family-specific motifs upon subgroup-specific features and coevolving protein domains. We propose the oldest tyrosine kinases, IRKL, SrcM, and Csk, originated from unicellular premetazoans and were coopted for complex multicellular functions. The increased frequency of oncogenic variants in more recent tyrosine kinases suggests that lineage-specific functionalities are selectively altered in human cancers.  相似文献   
102.
MCAK is a member of the kinesin-13 family of microtubule (MT)-depolymerizing kinesins. We show that the potent MT depolymerizer MCAK tracks (treadmills) with the tips of polymerizing MTs in living cells. Tip tracking of MCAK is inhibited by phosphorylation and is dependent on the extreme COOH-terminal tail of MCAK. Tip tracking is not essential for MCAK's MT-depolymerizing activity. We propose that tip tracking is a mechanism by which MCAK is preferentially localized to regions of the cell that modulate the plus ends of MTs.  相似文献   
103.
Colchicine is a known tubulin binding agent enabling necrosis in tumors. A novel tubulin-directed DO3A-colchicine conjugate and its Gd(III) complex were prepared from N-deacetylcolchicine, coupling alkaloid and polyaza-alicyclic functions via a peptide coupling methodology. The longitudinal proton relaxivity of the Gd(III) complex in water at 4.7 T is 2.86 mM−1 s−1 and a similar efficacy as colchicine towards ovarian carcinoma cells in vitro.  相似文献   
104.
105.
Chu HM  Yun M  Anderson DE  Sage H  Park HW  Endow SA 《The EMBO journal》2005,24(18):3214-3223
Kar3, a kinesin-14 motor of Saccharomyces cerevisiae required for mitosis and karyogamy, reportedly interacts with Cik1, a nonmotor protein, via its central, predicted coiled coil. Despite this, neither Kar3 nor Cik1 homodimers have been observed in vivo. Here we show that Kar3 is a dimer in vitro by analytical ultracentrifugation. The motor domains appear as paired particles by rotary-shadow electron microscopy (EM) and circular dichroism (CD) spectroscopy of the nonmotor region shows characteristics of helical structure, typical of coiled coils. Remarkably, the Kar3/Cik1 nonmotor region shows greater helicity by CD analysis and rotary-shadow EM reveals a stalk joined to one large or two smaller particles. The highly helical Kar3/Cik1 nonmotor region and visible stalk indicate that dimerization with Cik1 causes structural changes in Kar3. The Cik1 and Kar3 stalk regions preferentially associate with one another rather than forming homodimers. Kar3/Cik1 moves on microtubules at 2-2.4 microm min(-1), 2-5-fold faster than Kar3, and destabilizes microtubules at the lagging ends. Thus, structural changes in Kar3 upon dimerization with Cik1 alter the motor velocity and likely regulate Kar3 activity in vivo.  相似文献   
106.
The site-specific recombinase IntI1 found in class 1 integrons catalyzes the excision and integration of mobile gene cassettes, especially antibiotic resistance gene cassettes, with a site-specific recombination system. The integron integrase belongs to the tyrosine recombinase (phage integrase) family. The members of this family, exemplified by the lambda integrase, do not share extensive amino acid identities, but three invariant residues are found within two regions, designated box I and box II. Two conserved residues are arginines, one located in box I and one in box II, while the other conserved residue is a tyrosine located at the C terminus of box II. We have analyzed the properties of IntI1 variants carrying point mutations at the three conserved residues of the family in in vivo recombination and in vitro substrate binding. We have made four proteins with mutations of the conserved box I arginine (R146) and three mutants with changes of the box II arginine (R280); of these, MBP-IntI1(R146K) and MBP-IntI1(R280K) bind to the attI1 site in vitro, but only MBP-IntI1(R280K) is able to excise cassettes in vivo. However, the efficiency of recombination and DNA binding for MBP-IntI1(R280K) is lower than that obtained with the wild-type MBP-IntI1. We have also made two proteins with mutations of the tyrosine residue (Y312), and both mutant proteins are similar to the wild-type fusion protein in their DNA-binding capacity but are unable to catalyze in vivo recombination.Integrons are DNA elements that capture genes, especially antibiotic resistance genes, by a site-specific recombination system (32). The recombination system consists of a DNA integrase (Int) and two types of recombination sites, attI and attC (59-base element). The integrase gene (int) is located in the 5′ conserved segment of the integron structure (Fig. (Fig.1)1) and is a member of the tyrosine recombinase family (1, 4, 13, 23, 24). Three types of integrases, sharing around 50% identity among themselves, have been identified; they define integron classes 1, 2, and 3 (30). The 5′ conserved segment found in class 1 integrons also contains a promoter region responsible for the expression of inserted cassettes (11, 21) and the recombination site attI1 (31). The 3′ conserved segment of the class 1 integrons includes an ethidium bromide resistance determinant (qacEΔ1), a sulfonamide resistance gene (sulI), an open reading frame (ORF5) of unknown function, and further sequences that differ from one integron to another (5, 6, 28). The 3′ conserved segment of class 2 integrons includes transposition genes (20) while that of class 3 integrons has not yet been studied (2). The variable region, located between the two conserved segments, usually contains antibiotic resistance genes; In0 contains no inserted genes while In21 possesses eight cassettes with ten genes (or ORFs) in this region (5, 16). These genes are part of mobile cassettes which include a recombination site, attC, that differs from one gene to another (18, 33). Incoming genes must be associated with an attC to be recognized by the integron integrase and are preferentially inserted at the recombination site attI1 (11). Cassettes are excised as circular intermediates and integrated at core sites by the action of the integrase (810). The core site, defined as GTTRRRY, makes up the 3′ end of attI1 and attC, with the crossover taking place between the G and the first T (19). Antibiotic selection pressure can reveal cassette rearrangements in which a given resistance is nearest the promoter and thus most strongly expressed (10). Open in a separate windowFIG. 1General structure of class 1 integrons. Cassettes are inserted in the integron variable region by a site-specific recombination mechanism. The attI1 site is shown by a black circle, core sites are represented by ovals, the attC site is indicated by a black rectangle, and promoters are denoted by P. intIl, integrase gene; qacEΔ1, antiseptic resistance gene; sulI, sulfonamide resistance gene; orf5, gene of unknown function.Site-specific recombination, unlike homologous recombination, is characterized by relatively short, specific DNA sequences and requires only limited homology of the recombining partners (12). Site-specific recombination is an entirely conservative process since all DNA strands that are broken (two per exchange site) are rejoined in a process that involves neither ATP nor DNA synthesis. Homology alignments of site-specific recombinases assign them to two families: the resolvase family, named after the TnpR proteins encoded by the transposons γδ and Tn3, and the integrase family. The integrase family includes over 140 members to date, but they are highly diversified proteins (13, 23). Members of this family, which include the well-studied λ integrase, recombine DNA duplexes by executing two consecutive strand breakage and rejoining steps and a topoisomerization of the reactants. The first pair of exchanges form a four-way Holliday junction and the second pair resolve the junction to complete the recombination. The integrase nucleophile is a conserved tyrosine that becomes associated with a phosphate group on DNA. The cleavage sites on each DNA duplex are separated by 6 to 8 bp with a 5′ stagger, and the tyrosine joins to the 3′ phosphate (17).The initial definition of the integrase family was based on comparisons of seven sequences, and three invariant residues were identified: an HXXR cluster and a Y residue (4). Alignment of 28 sequences identified a fourth invariant position, occupied by an arginine residue (1). These four conserved residues are found in two boxes located in the second half of the protein. A recent analysis has shown that the conserved histidine is present in 136 of the 147 members (93%); this residue is then not conserved in all members of the family (13). Another recent analysis has identified three patches of residues located around box I, which seem to be important in the secondary structure of these proteins (23). In this study, we analyzed the properties of several mutants of the conserved residues R146, R280, and Y312 of the integron integrase IntI1 in in vivo recombination and in vitro substrate binding.

Construction of plasmids overexpressing mutant MBP-IntI1 fusion proteins.

The plasmids encoding various mutants of MBP-IntI1 were constructed by PCR using pLQ369 (50 ng) as a template (15). Two primer pairs, designed with the OLIGO software package (version 4.1; National Biosciences, Plymouth, Minn.), were used to construct each set of mutants. The R146 mutants were constructed with an XcmI-BamHI primer pair [IntI1(R146)-XcmI, 5′-TTCACCAGCTTCTGTATGGAACGGGCATG(A/G)(A/T)AATCAG-3′; IntI1(R146)-BamHI, 5′-CCGGATCCCTACCTCTCACT-3′], the R280 mutants were constructed with an NruI-XmnI primer pair [IntI1(R280)-NruI, 5′-AGCCGTCGCGAACGAGTGC(C/T)(C/T)GAGGG-3′; IntI1(R280)-XmnI, 5′-ACCCCTAATGAAGTGGTTCGTATCC-3′], and the Y312 mutants were constructed with a AatII-ScaI primer pair [IntI1(Y312)-AatII, 5′-ATTCCGACGTCTCTACTACGATGATTT(C/T)CACGC-3′; pLQ369-ScaI, 5′-ATGCTTTTCTGTGACTGGTG-3′] (restriction sites within primer sequences are underlined). PCR conditions were 10 min at 94°C, three cycles consisting of 45 s at 94°C, 45 s at 47°C, and 90 s at 72°C, 30 cycles consisting of 45 s at 94°C, 45 s at 60°C (50°C for Y312 mutants), and 90 s at 72°C, and a final elongation step of 10 min at 72°C. The XcmI, NruI, and AatII primers were degenerate in one or two positions, so that a single primer could give all mutants. Mutant PCR fragments were digested and cloned directly into pLQ369 digested with the same enzymes, except for the R146 mutant fragments that were subcloned into pLQ364 at first. New mutant PCR fragments were then amplified on these subclones, using IntI1(R146)-BamHI and IntI1(R280)-XmnI primers. These mutant PCR fragments were cleaved with BamHI and XmnI, and the resulting fragments were cloned into pLQ369. This avoids the necessity of partial digestion of pLQ369 with XcmI. Mutant clones were digested with restriction endonucleases and sequenced to determine the mutation.

In vivo recombination.

Mutant MBP-IntI1 clones were introduced into Escherichia coli TB1 {F′ araΔ(lac-proAB) rpsL (Strr) [φ80dlacΔ(lacZ)M15] hsdR(rKmK)} containing pLQ428 by transformation (Fig. (Fig.22 and Table Table1).1). E. coli TB1 cells containing pLQ428 and one of the MBP-IntI1 mutants were grown at 37°C for 3 h in Luria-Bertani medium. Excision of the aacA1-ORFG and/or ORFH cassettes was induced by the overexpression of the malE-intI1 gene by using 0.3 mM isopropyl-β-d-thiogalactopyranoside (IPTG; Sigma Chemical Co.) and by incubation at 37°C for another 3 h. Cell culture was done in the presence of 50 μg of ampicillin per ml, 15 μg of amikacin per ml, and 50 μg of chloramphenicol per ml. Plasmid DNA was then prepared from 5-ml cultures with the Perfect Prep DNA extraction kit (Mandel Corporation). In order to determine the capacity of mutant MBP-IntI1 proteins to excise aacA1-ORFG and/or ORFH cassettes of In21, we used PCR primers pACYC184-5′ (5′-TGTAGCACCTGAAGTCAGCC-3′) and pACYC184-3′ (5′-ATACCCACGCCGAAACAAG-3′) (Fig. (Fig.2,2, primers 1 and 2) to detect the reduction of pLQ428 length. PCR conditions were 10 min at 94°C, 30 cycles consisting of 1 min at 94°C, 1 min at 60°C, and 5 min at 72°C, and a final elongation step of 10 min at 72°C. A major PCR fragment can be seen in each lane containing a DNA preparation from a mutant clone (Fig. (Fig.3,3, lanes 2 to 9). This band is 2,499 bp long and, as determined by restriction enzyme digestions, represents the pLQ428 clone without any cassette excision (data not shown). This band is also observed in the negative control, which is the pMAL-c2 vector without any gene fused to malE (Fig. (Fig.3,3, lane 12). Open in a separate windowFIG. 2Representation of plasmids used in this study. The positions of the three invariant residues of the integrase family are indicated, along with restriction sites used to construct mutant proteins. Core sites are represented by black circles, and attCs are shown by white boxes. The numbered arrows represent the PCR primers used to detect excision events, pACYC184-5′ (1) and pACYC184-3′ (2). bla, gene encoding β-lactamase; cat, gene encoding chloramphenicol acetyltransferase; intIl, gene encoding the integron integrase (IntI1); malE, gene encoding the maltose binding protein (MBP); ori, origin of replication; Ptac, tac promoter; Ptet, tetracycline promoter. Only relevant restriction sites are indicated.

TABLE 1

Plasmids used in this study
PlasmidCharacteristic(s)aReference or source
pLQ3632,190-bp EcoRI-HincII fragment of pLQ161 cloned in pLQ402 (Apr)16
pLQ3641,027-bp NcoI-BamHI PCR fragment amplified on pLQ860 and cloned in pET-3d (Apr)This study
pLQ3691,019-bp NdeI-BamHI PCR fragment modified to create a blunt-end 5′-ATG and cloned in pMAL-c2 cut with XmnI-BamHI (Apr)15
pLQ376pLQ369 MBP-IntI1(R146K) (Apr)This study
pLQ377pLQ369 MBP-IntI1(R146E) (Apr)This study
pLQ378pLQ369 MBP-IntI1(R146I) (Apr)This study
pLQ379pLQ369 MBP-IntI1(R146V) (Apr)This study
pLQ388pLQ369 MBP-IntI1(R280G) (Apr)This study
pLQ390pLQ369 MBP-IntI1(R280E) (Apr)This study
pLQ391pLQ369 MBP-IntI1(R280K) (Apr)This study
pLQ393pLQ369 MBP-IntI1(Y312S) (Apr)This study
pLQ394pLQ369 MBP-IntI1(Y312F) (Apr)This study
pLQ4282,133-bp EcoRI (filled in)-BglII fragment of pLQ363 cloned in pACYC184 cut with EcoRV-BamHI (Akr Cmr)This study
pLQ8602,900-bp BamHI fragment of pVS1 cloned in pTZ19R (Apr Sulr)5
Open in a separate windowaAkr, Apr, and Cmr, resistance to amikacin, ampicillin, and chloramphenicol. Open in a separate windowFIG. 3Electrophoresis of PCR products obtained with the pACYC184 primer pair and 100 ng of DNA preparations from overexpressed cultures on a 1% agarose gel. Lane 1, 1-kb DNA ladder (Gibco BRL); lane 2, DNA preparation of pLQ428-pLQ377 (R146E); lane 3, pLQ428-pLQ378 (R146I); lane 4, pLQ428-pLQ376 (R146K); lane 5, pLQ428-pLQ379 (R146V); lane 6, pLQ428-pLQ390 (R280E); lane 7, pLQ428-pLQ388 (R280G); lane 8, pLQ428-pLQ391 (R280K); lane 9, pLQ428-pLQ394 (Y312F); lane 10, pLQ428-pLQ393 (Y312S); lane 11, pLQ428-pLQ369 (wild type); lane 12, pLQ428-pMAL-c2 (MBP).The 2,499-bp PCR product was not obtained in the reaction containing the wild-type MBP-IntI1-expressing clone pLQ369 (Fig. (Fig.3,3, lane 11), indicating that there were no remaining full-length pLQ428 molecules. This shows that the wild-type fusion protein is very efficient in site-specific recombination and that all pLQ428 clones have undergone an excision of one or both cassettes. In this PCR, we observed two major bands of 1,341 and 889 bp. The 1,341-bp PCR product was digested with restriction enzymes to show that it represents a pLQ428 clone which has lost the aacA1-ORFG cassette (data not shown). The 889-bp band was also digested with restriction enzymes to show that it represents a pLQ428 clone which has lost both aacA1-ORFG and ORFH cassettes (data not shown). These two PCR products are also observed in the reaction containing the mutant clone pLQ391, which expresses the MBP-IntI1(R280K) fusion protein. This mutant protein is, however, less efficient than the wild-type protein, as seen by the intensity of the PCR products (Fig. (Fig.3,3, lane 8). We were not able to detect a PCR product of 2,047 bp, corresponding to the excision of the ORFH cassette alone; this is not surprising since this event has been shown in another study to be rare (16). It is possible to observe another band in pLQ428-pLQ391 (R280K) and pLQ428-pLQ369 (wild type) PCRs (Fig. (Fig.3,3, lanes 8 and 11); this PCR product is 1,100 bp long and probably represents a recombination event at a secondary site. Restriction enzyme digestions were done on this product, but we were unable to identify its origin. This product results from an event mediated by the integron integrase since it is seen only in reactions containing active proteins. An 1,800-bp PCR band is also present in the negative control and in all PCRs containing a mutant clone. This product appears to be nonspecific, and the fact that it is not seen in the PCR containing the pLQ428-pLQ369 (wild-type) clones probably results from the PCR being more favorable to smaller PCR products.

In vitro substrate binding.

The experiments described above demonstrate that only one of our mutants of IntI1 protein is able to catalyze in vivo recombination. Can all mutant proteins recognize and bind to the IntI1 recombination site in a manner similar to the wild-type protein? To investigate this question, we used purified fusion proteins and a gel retardation assay with the complete attI1 site (5′ site) of the integron. MBP-IntI1 fusion proteins were purified as suggested by New England Biolabs. The concentration of the purified fusion protein was determined by using the Bradford protein assay (Bio-Rad). The protein solution was then made 20% in glycerol and stored at −80°C. The purity of MBP-IntI1 was evaluated as >90% by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (data not shown). Binding reactions were done with labeled 5′-site DNA fragments (20,000 cpm, 0.25 pmol), incubated with different concentrations of MBP-IntI1 in a 10-μl volume containing 10 mM HEPES (K+, pH 8.0), 60 mM KCl, 4 mM MgCl2, 100 μM EDTA (pH 8.0), 100 μg of bovine serum albumin per ml, 250 μM dithiothreitol, 100 ng of poly(dI-dC), and 10% glycerol. Reaction mixtures were incubated at room temperature for 15 min prior to electrophoresis through 4 or 5% prerun, nondenaturing polyacrylamide gels buffered with 0.5× Tris-borate-EDTA. Dried gels were subjected to autoradiography. The wild-type fusion protein and native IntI1 were shown to lead to the same four distinct complexes (I, II, III, and IV) with this DNA substrate (Fig. (Fig.4)4) (15). These complexes represent the binding of four IntI1 molecules to four different sites in the attI1 site (15). Figure Figure44 shows results obtained with nine mutants of the MBP-IntI1 fusion protein. We observed that MBP-IntI1(R146E), MBP-IntI1(R146I), and MBP-IntI1(R146V) lost their ability to bind to the attI1 site, as no complexes are seen in the gel retardation experiment (Fig. (Fig.4A).4A). However, MBP-IntI1(R146K) formed four IntI1-DNA complexes with the 5′ site DNA fragment. The band pattern and the intensity observed with this mutant protein are similar to those observed with the wild-type protein, suggesting that MBP-IntI1(R146K) and MBP-IntI1 bind DNA with similar affinities. Open in a separate windowFIG. 4Binding of mutant MBP-IntI1 fusion proteins purified from E. coli TB1 to the 5′-site DNA fragment containing the complete attI1 site of the In2 integron (from nucleotide −96 to nucleotide +71, relative to the G residue of the core site as position 0). (A) MBP-IntI1(R146) mutants; (B) MBP-IntI1(R280) mutants; (C) MBP-IntI1(Y312) mutants. A purified labeled fragment was incubated with different concentrations of mutant fusion proteins. Free DNA (F) and protein-DNA complexes (I, II, III, and IV) were separated on 4 or 5% polyacrylamide gels and are indicated by arrows. Lanes 1, free DNA; lanes 2 through 7, purified fusion protein at 250, 375, 500, 12.5, 37.5, and 62.5 nM, respectively. The wild-type (WT) lanes in panel C were from a separate gel.Competition with a specific fragment with a 30-fold excess of unlabeled DNA competed away all four complexes, while a 100-fold excess of a nonspecific unlabeled DNA fragment did not compete away any complexes, indicating their specificity (data not shown) (15). We observed that MBP-IntI1(R280G) and MBP-IntI1(R280E) lost their ability to bind the 5′-site DNA fragment, while the MBP-IntI1(R280K) could still bind the attI1 site (Fig. (Fig.4B).4B). However, the band pattern obtained with this mutant protein is weaker than that obtained with the wild-type integrase. For example, at a protein concentration of 250 nM MBP-IntI1(R280K) (lane 2), we observed the formation of complexes I, II, and III, with a stronger intensity for the fastest-migrating complexes, while the intensity of the fourth complex was very weak. At the same concentration of the wild-type protein, we observed the formation of all four complexes, with a stronger intensity for the slowest-migrating complexes and no unbound DNA. These results show that MBP-IntI1(R280K) binds the attI1 site with a lower affinity than the wild-type fusion protein. As shown in Fig. Fig.4C,4C, both MBP-IntI1(Y312F) and MBP-IntI1(Y312S) lead to the formation of four complexes that migrate similarity to those obtained with wild-type MBP-IntI1, as judged by the gel migration of these complexes. The band pattern observed shows that the binding affinity of these mutant proteins is the same as or even better than that of the wild-type protein.

Relationships with other members of the family.

We found that MBP-IntI1 recombinase in which Arg-146 has been changed to lysine [MBP-IntI1(R146K)] by PCR mutagenesis cannot excise cassettes but can bind to the attI1 site with the same efficiency as the wild-type fusion protein. However, MBP-IntI1(R146I), MBP-IntI1(R146E), and MBP-IntI1(R146V) mutant proteins have completely lost both phenotypes. These findings are different from those for other members of the family. The only mutant protein of the lambda integrase at this residue [λ(R212Q)] binds the core site partially and is not able to catalyze in vivo or in vitro recombination (22). Mutants of the Cre recombinase with a change at this residue [Cre(R173K)] bind DNA as well as the wild-type protein but cannot catalyze in vivo or in vitro recombination (1). Mutants of Flp [Flp(R191K) and Flp(R191E)] bind FRT recombination sites as well as the wild-type protein but cannot carry out in vivo or in vitro recombination, except for the Flp(R191K) protein, which has shown slight activity in in vivo recombination (Table (Table2)2) (7, 14, 25). Therefore, the Cre(R173K) and Flp(R146K) mutants have the same phenotype as the MBP-IntI1(R146K) protein. However, the Flp(R191E) mutant protein shows efficient DNA binding while MBP-IntI1(R146E) does not bind to the attI1 site. We interpret these results according to the charge of the Arg-146 residue. The positively charged side chain of this residue makes contact with the DNA, which is negatively charged. This contact is probably important for the good conformation of the protein molecule in positioning the tyrosine residue to perform recombination. When this residue is exchanged for a lysine, DNA contacts are still able to take place because of the charge of the residue, but the side chain is smaller and the lysine is probably not able to position the tyrosine to catalyze recombination. We think that the charge of this residue is very important in the formation of DNA-protein complexes in the integron system, since all other MBP-IntI1 mutants tested are unable to bind DNA. This observation differs from those for Flp, because even when the wild-type residue was replaced by a negatively charged one, it could still bind DNA as well as the wild-type protein (Table (Table2).2).

TABLE 2

Mutational analysis of IntI1 and corresponding residues of other recombinases from the Int family
RecombinaseMutationDNA bindingRecombinationReference(s)
λIntR212QYesaNo22
λIntY342FYesNo22, 26
FlpR191EYesNo7
FlpR191KYesYes7, 14
FlpR308GYesNo27
FlpR308KYesYesa27
FlpY343FYesNo29
FlpY343SYesNo29
CreR173KYesNo1
P2R272KNDbNo23
XerCY275FYesNo3
XerDY279FYesNo3
IntI1R146ENoNoThis study
IntI1R146INoNoThis study
IntI1R146KYesNoThis study
IntI1R146VNoNoThis study
IntI1R280ENoNoThis study
IntI1R280GNoNoThis study
IntI1R280KYesYesaThis study
IntI1Y312FYesNoThis study
IntI1Y312SYesNoThis study
Open in a separate windowaLess efficient than the wild-type protein. bND, not determined. We have also made proteins with mutations at position 280; these were MBP-IntI1(R280E), MBP-IntI1(R280G), and MBP-IntI1(R280K). We found that the MBP-IntI1(R280K) mutant protein binds the attI1 site and excises integron cassettes with a lower efficiency than the wild-type MBP-IntI1, while MBP-IntI1(R280E) and MBP-IntI1(R280G) have completely lost both phenotypes. The Flp(R308K) mutant protein has been shown to bind DNA as well as the wild-type protein, but it recombines DNA with a lower efficiency than wild-type Flp (27). Another mutant protein of Flp [Flp(R308G)] has also been shown to bind DNA as well as the wild-type protein, but it was unable to catalyze in vivo or in vitro recombination (27). These results show that Flp(R308K) and MBP-IntI1(R280K) act similarly but that the other Flp mutant [Flp(R308G)] can bind DNA while the MBP-IntI1 mutant [MBP-IntI1(R280G)] cannot (Table (Table2).2). We also think that the positive charge of this residue is important for the binding of the recombinase to DNA, but Arg-280 does not seem to be implicated in the positioning of the tyrosine residue, since the MBP-IntI1(R280K) mutant protein can perform recombination.We found that MBP-IntI1(Y312S) and MBP-IntI1(Y312F) mutant proteins bind the attI1 site with the same efficiency as the wild-type protein but are not able to catalyze in vivo recombination. As expected, these results are the same as those obtained with the lambda integrase [λ(Y342F)], the XerC and XerD recombinases [XerC(Y275F) and XerD(Y279F)], and the Flp recombinases [Flp(Y343S) and Flp(Y343F)] (Table (Table2)2) (3, 22, 26, 29). The loss of the catalytic activity of the MBP-IntI1(Y312F) mutant protein is not surprising, since the hydroxyl group of the tyrosine, which is responsible for the nucleophilic attack of the DNA at the recombination site, is not present on the phenylalanine residue. The phenotype of MBP-IntI1(Y312S) indicates that the conformation of the tyrosine residue is important for the good activity of the recombinase, because even if the serine residue has a hydroxyl group, it is not able to catalyze recombination. These results indicate that the integron integrase IntI1 uses the hydroxyl group of the conserved tyrosine (Y312) to catalyze site-specific recombination, like other members of the family. However, in vitro recombination using this mutant protein needs to be done to confirm this.These results of point mutations show that mutations of the conserved arginines by nonpositively charged residues abolish substrate recognition, unlike the corresponding mutants of other members of the family. However, further mutational analysis, such as of residues around and in patch III, would be interesting, since only integron integrases contain more residues in this region than other members of the family (23). In vitro recombination assays with purified mutant proteins also need to be done in order to study thoroughly the mechanism of site-specific recombination in integrons.  相似文献   
107.
Anthesis and pollen dispersal of H. lanatus and F. rubra were related to climate factors during a field investigation on Schiermonnikoog. In both species the anthesis showed a diurnal periodicity and could be connected to air temperature, relative air humidity, and light intensity. In the pollen release of both species no daily periodicity was found. This pollen release appears to be dependent on neither the temperature nor the light intensity. It could be related to only one of the measured climate factors, viz. the relative air humidity. It was discussed that anthesis is an active process based on physiological and ecological properties of the plant, whereas the pollen dispersal is only a mechanical process.  相似文献   
108.
Previous results from this laboratory have shown that very low chronic doses of gamma radiation can stimulate proliferation of the Cyanobacterium Synechococcus lividus. This modification of cell proliferation occurred during the first doubling. In this paper, we have compared the metabolism of cells cultivated in a normal environment or under chronic irradiation. Incubation of the cells in a new medium induced a high superoxide dismutase (EC 1.15.1.1, SOD) activity at the 18th hour and a degradation of phycocyanin, thus demonstrating that cells were submitted to a photooxidative stress. This increase in superoxide dismutase activity was followed by concomittant peaks of glutathione reductase (EC 1.6.4.2, GR) and glucose-6-phosphate dehydrogenase (EC 1.1.1.49, G6P-DH) at the 24th hour. Irradiated cultures at a dose of 53.5 mGray/year show an earlier and higher peak of SOD, GR, and G6P-DH. In a second stage, cultures showed an earlier onset of photosynthesis under irradiation, as evidenced by an increase in pigment content and an enhancement of glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.13, GAP-DH). These results show that the radiostimulation is related to the activation of enzymes protecting against peroxides that were induced under oxidative circumstances and to the activation of a glucose catabolism via the oxidative pentose phosphate pathway.Abbreviations mGy milli-Gray - SOD superoxide dismutase - G6P-DH glucose-6-phosphate dehydrogenase - GAP-DH glycer-aldehyde-3-phosphate dehydrogenase - GSSG oxidized glutathione  相似文献   
109.
The ecology, abundance and diversity of galatheoid squat lobsters make them an ideal group to study deep-sea diversification processes. Here, we reconstructed the evolutionary and biogeographic history of Leiogalathea, a genus of circum-tropical deep-sea squat lobsters, in order to compare patterns and processes that have affected shallow-water and deep-sea squat lobster species. We first built a multilocus phylogeny and a calibrated species tree with a relaxed clock using StarBEAST2 to reconstruct evolutionary relationships and divergence times among Leiogalathea species. We used BioGeoBEARS and a DEC model, implemented in RevBayes, to reconstruct ancestral distribution ranges and the biogeographic history of the genus. Our results showed that Leiogalathea is monophyletic and comprises four main lineages; morphological homogeneity is common within and between clades, except in one; the reconstructed ancestral range of the genus is in the Atlantic and Indian oceans (Tethys). They also revealed the divergence of the Atlantic species around 25 million years ago (Ma), intense cladogenesis 15–25 Ma and low levels of speciation over the last 5 million years (Myr). The four Leiogalathea lineages showed similar patterns of speciation: allopatric speciation followed by range expansion and subsequent stasis. Leiogalathea started diversifying during the Oligocene, likely in the Tethyan. The Atlantic lineage then split from its Indo-Pacific sister group due to vicariance driven by closure of the Tethys Seaway. The Atlantic lineage is less speciose compared with the Indo-Pacific lineages, with the Tropical Southwestern Pacific being the current centre of diversity. Leiogalathea diversification coincided with cladogenetic peaks in shallow-water genera, indicating that historical biogeographic events similarly shaped the diversification and distribution of both deep-sea and shallow-water squat lobsters.  相似文献   
110.
Among the recent introductions of alien insects in Europe, the polyphagous western conifer seed bug, Leptoglossus occidentalis Heidemann (Heteroptera; Coreidae) can seriously be regarded as a major threat for all the European conifer forests. In the current study combining laboratory and field experimentations, we characterized first bug damage by developing specific damage categories on seeds of different conifer species by the use of X-ray. Secondly, we investigated the impact of the invasive bug on key conifer species used for afforestation in Western and Central Europe. For this purpose, we performed germination tests on predated seeds which revealed that even light damage (consumption of <1/3 of the whole seed content) strongly reduced the germination capability of the seed. We also compared the impact of feeding on the proportion of filled seeds. Second year cones of Pinus sylvestris and Pinus nigra have been enclosed and offered to different life stages (nymphs and adults) and the results showed a significant reduction of filled seeds whatever the life stage. In field, we annually surveyed the bug seed damage for six different conifer species planted in southwestern French seed orchards. Taking into account the economic value of improved seeds in seed orchards, economic impact of bug damage was important although never exceeded 25 %. Two natural or semi-natural alpine pine stands were also surveyed and appeared to be highly affected by the bug (up to 70 % of damaged seeds). Therefore, bug damage could also be considered as a serious threat for seed production in natural stands.  相似文献   
设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号