首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
AEM. Adams  D. Botstein 《Genetics》1989,121(4):675-683
A gene whose product is likely to interact with yeast actin was identified by the isolation of pseudorevertants carrying dominant suppressors of the temperature-sensitive (Ts) act1-1 mutation. Of 30 independent revertants analyzed, 29 were found to carry extragenic suppressor mutations and of these, 24/24 tested were found to be linked to each other. This linkage group identifies a new gene SAC6, whose product, by several genetic criteria, is likely to interact intimately with actin. First, although act1-1 sac6 strains are temperature-independent (Ts+), 4/17 sac6 mutant alleles tested are Ts in an ACT1+ background. Moreover, four Ts+ pseudorevertants of these ACT1+ sac6 mutants carry suppressor mutations in ACT1; significantly, three of these are again Ts in a SAC6+ background, and are most likely new act1 mutant alleles. Thus, mutations in ACT1 and SAC6 can suppress each other's defects. Second, sac6 mutations can suppress the Ts defects of the act1-1 and act1-2, but not act1-4, mutations. This allele specificity indicates the sac6 mutations do not suppress by simply bypassing the function of actin at high temperature. Third, act1-4 sac6 strains have a growth defect greater than that due to either of the single mutations alone, again suggesting an interaction between the two proteins. The mutant sac6 gene was cloned on the basis of dominant suppression from an act1-1 sac6 mutant library, and was then mapped to chromosome IV, less than 2 cM from ARO1.  相似文献   

2.
3.
Spontaneous Unstable UNC-22 IV Mutations in C. ELEGANS Var. Bergerac   总被引:19,自引:2,他引:19  
This paper describes a mutator system in the nematode Caenorhabditis elegans var. Bergerac for the gene unc-22. Of nine C. elegans and two C. briggsae strains tested only the Bergerac BO strain yielded mutant animals at a high frequency and the unc-22 IV gene is a preferred mutational target. The forward spontaneous mutation frequency at the unc-22 locus in Bergerac BO is about 1 x 10-4 , and most of these spontaneous unc-22 mutations revert at frequencies between 2 x 10-3 and 2 x 10 -4. Both the forward mutation frequency and the reversion frequency are sensitive to genetic background. Spontaneous unc-22 mutations derived in a Bergerac background and placed in a primarily Bristol background revert at frequencies of <10-6. When reintroduced into a Bergerac/Bristol hybrid background the mutations once again become unstable.

The mutator activity could not be localized to a discrete site in the Bergerac genome. Nor did mutator activity require the Bergerac unc-22 gene as a target since the Bristol unc-22 homolog placed in a Bergerac background also showed high mutation frequency. Intragenic mapping of two spontaneous unc-22 alleles, st136 and st137, place both mutations in the central region of the known unc-22 map. However, these mutations probably recombine with one another, suggesting that the unstable mutations can occur in more than one site in unc-22. Examination of the phenotypic effect of these mutations on muscle structure indicates that they are less severe in their effect than a known amber allele. We suggest that this mutator system is polygenic and dispersed over the nematode genome and could represent activity of the transposable element Tc1.

  相似文献   

4.
A Visible Allele of the Muscle Gene sup-10 X of C. ELEGANS   总被引:3,自引:2,他引:1       下载免费PDF全文
In this paper, we extend our previous analyses of a set of genes in Caenorhabditis elegans that are involved in muscle structure and function: unc-93 III, sup-9 II, sup-10 X and sup-11 I. We describe an unusual, visible allele of sup-10, examine how this allele interacts genetically with mutations in other genes of this set and propose that the wild-type products of the unc-93 and sup-10 loci may be components of a protein complex. We also describe a new gene of this set, sup-18 III, and the interaction of sup-18 alleles with mutations in the other genes.  相似文献   

5.
Ninety-five mutants of the nematode Caenorhabditis elegans altered in the cell lineages of the vulva have been isolated on the basis of their displaying one of two phenotypes, Vulvaless or Multivulva. In Vulvaless mutants, which define 12 genes, no vulva is present. In Multivulva mutants, which define ten genes, one or more supernumerary vulva-like protrusions are located along the ventral side of the animal. A single recessive mutation is responsible for the phenotypes of most, but not all, of these strains. Fifteen of these 22 genes are represented by multiple alleles. We have shown by a variety of genetic criteria that mutations that result in a Vulvaless or Multivulva phenotype in six of the 22 genes most likely eliminate gene function. In addition, Vulvaless or Multivulva mutations in seven of the other genes most likely result in a partial reduction of gene function; the absence of the activity of any of these genes probably results in lethality or sterility. Our results suggest that we may have identified most, or all, genes of these two classes.  相似文献   

6.
Evidence exists, particularly in the welfare literature of nonhuman animals on the farm, that the interaction between nonhuman animals and the personnel who care for them can have a strong effect on the animals' behavior, productivity, and welfare. Among species commonly used for biomedical research, mice appear to be the least-preferred species in animal care facilities. A review of the literature and observations of animal care staff interacting with mice indicated that the following factors may influence this: their small size, their particular behavioral characteristics, and husbandry constraints (such as housing in ventilated racks). In addition, this study questioned whether animal care personnel have a different perception of genetically engineered animals and whether this, in turn, has an effect on their interactions with these animals. The ability to carefully observe an animal's behavior is key in carrying out an animal-wellness assessment and in minimizing pain and distress. Attention to human-animal interactions in the research setting represents an opportunity for refinement for large numbers of animals and potentially for reduction of animal use.  相似文献   

7.
Golgi anti-apoptotic proteins (GAAPs) are multitransmembrane proteins that are expressed in the Golgi apparatus and are able to homo-oligomerize. They are highly conserved throughout eukaryotes and are present in some prokaryotes and orthopoxviruses. Within eukaryotes, GAAPs regulate the Ca2+ content of intracellular stores, inhibit apoptosis, and promote cell adhesion and migration. Data presented here demonstrate that purified viral GAAPs (vGAAPs) and human Bax inhibitor 1 form ion channels and that vGAAP from camelpox virus is selective for cations. Mutagenesis of vGAAP, including some residues conserved in the recently solved structure of a related bacterial protein, BsYetJ, altered the conductance (E207Q and D219N) and ion selectivity (E207Q) of the channel. Mutation of residue Glu-207 or -178 reduced the effects of GAAP on cell migration and adhesion without affecting protection from apoptosis. In contrast, mutation of Asp-219 abrogated the anti-apoptotic activity of GAAP but not its effects on cell migration and adhesion. These results demonstrate that GAAPs are ion channels and define residues that contribute to the ion-conducting pore and affect apoptosis, cell adhesion, and migration independently.  相似文献   

8.
Meredith Kusch  R. S. Edgar 《Genetics》1986,113(3):621-639
In Caenorhabditis elegans, four loci (sqt-1, sqt-2, sqt-3 and rol-8) in which mutations affect body shape and cuticle morphology have unusual genetic properties. Mutant alleles of sqt-1 can interact to produce animals with a variety of mutant phenotypes: left roller, right roller, dumpy and long. At least three mutant phenotypes are specified by mutations in the sqt-3 locus. Most alleles at these loci are either dominant or cryptic dominant (i.e., are dominant only in certain genetic backgrounds). Most alleles of these loci exhibit codominance. Two putative null alleles of the sqt-1 locus produce a wild-type phenotype. Many alleles of these genes demonstrate unusual intergenic interactions that are not the result of simple epistasis: animals doubly heterozygous for mutations at two loci often display unexpected and unpredictable phenotypes. We suggest that these genetic properties might be expected of genes, such as the collagen genes, the products of which interact to form the animal's cuticle, and which are member genes of a gene family.  相似文献   

9.
Despite considerable progress understanding genes that affect the HDL particle, its function, and cholesterol content, genes identified to date explain only a small percentage of the genetic variation. We used N-ethyl-N-nitrosourea mutagenesis in mice to discover novel genes that affect HDL cholesterol levels. Two mutant lines (Hlb218 and Hlb320) with low HDL cholesterol levels were established. Causal mutations in these lines were mapped using linkage analysis: for line Hlb218 within a 12 Mbp region on Chr 10; and for line Hlb320 within a 21 Mbp region on Chr 7. High-throughput sequencing of Hlb218 liver RNA identified a mutation in Pla2g12b. The transition of G to A leads to a cysteine to tyrosine change and most likely causes a loss of a disulfide bridge. Microarray analysis of Hlb320 liver RNA showed a 7-fold downregulation of Hpn; sequencing identified a mutation in the 3' splice site of exon 8. Northern blot confirmed lower mRNA expression level in Hlb320 and did not show a difference in splicing, suggesting that the mutation only affects the splicing rate. In addition to affecting HDL cholesterol, the mutated genes also lead to reduction in serum non-HDL cholesterol and triglyceride levels. Despite low HDL cholesterol levels, the mice from both mutant lines show similar atherosclerotic lesion sizes compared to control mice. These new mutant mouse models are valuable tools to further study the role of these genes, their affect on HDL cholesterol levels, and metabolism.  相似文献   

10.
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   

11.
The interactions of the herpes simplex virus processivity factor UL42 with the catalytic subunit of the viral polymerase (Pol) and DNA are critical for viral DNA replication. Previous studies, including one showing that substitution of glutamine residue 282 with arginine (Q282R) results in an increase of DNA binding in vitro, have indicated that the positively charged back surface of UL42 interacts with DNA. To investigate the biological consequences of increased DNA binding by UL42 mutations, we constructed two additional UL42 mutants, including one with a double substitution of alanine for aspartic acid residues (D270A/D271A) and a triple mutant with the D270A/D271A and Q282R substitutions. These UL42 mutants exhibited increased and prolonged DNA binding without an effect on binding to a peptide corresponding to the C terminus of Pol. Plasmids expressing any of the three UL42 mutants with an increased positive charge on the back surface of UL42 were qualitatively competent for complementation of growth and DNA replication of a UL42 null mutant on Vero cells. We then engineered viruses expressing these mutant proteins. The UL42 mutants were more resistant to detergent extraction than wild-type UL42, suggesting that they are more tightly associated with DNA in infected cells. All three UL42 mutants formed smaller plaques on Vero cells and replicated to reduced yields compared with results for a control virus expressing wild-type UL42. Moreover, mutants with double and triple mutations, which contain D270A/D271A mutations, exhibited increased mutation frequencies, and mutants containing the Q282R mutation exhibited elevated ratios of virion DNA copies per PFU. These results suggest that herpes simplex virus has evolved so that UL42 neither binds DNA too tightly nor too weakly to optimize virus production and replication fidelity.Processivity factors of DNA polymerases promote long-chain DNA synthesis by preventing dissociation of the DNA polymerase from the primer/template. Processivity factors also can influence DNA replication fidelity, as indicated by numerous in vivo and in vitro studies (1-3, 5, 6, 11, 12, 18, 28, 36). A major class of processivity factors known as “sliding clamps” includes proliferating cell nuclear antigen (PCNA) of eukaryotic cells (23) and gp45 of T4 bacteriophage (27). Sliding clamps are homodimers or homotrimers that encircle DNA and interact with the catalytic subunits (Pols) of their cognate DNA polymerases to promote processive DNA synthesis.A second class of processivity factors includes those encoded by herpesviruses and is exemplified by herpes simplex virus (HSV) UL42. UL42 forms a heterodimer with the HSV Pol. Both subunits are essential for production of infectious virus and for viral DNA replication (20, 26). UL42 can stimulate long-chain DNA synthesis by Pol, and template challenge experiments established that this stimulation is due to increased processivity (15). In addition to its interaction with Pol, which is mediated by the C terminus of Pol, UL42 also binds DNA directly with high affinity (14, 15, 30, 37). This mode of DNA binding differs from that of sliding clamps, which do not form high-affinity direct interactions with DNA (13) but must be loaded onto DNA with the aid of ATP-dependent clamp loaders for their normal functioning (16). Nevertheless, the structure of UL42 is very similar to a monomer of the sliding clamp PCNA (39). Like other processivity factors, UL42 also plays a role in maintaining DNA replication fidelity both in vivo and in vitro (5, 18).The “back face” (opposite face to the side that binds Pol) of a UL42 molecule contains several positively charged residues. By titrating the effects of cations on UL42 DNA binding, it was determined that charge-charge interactions are involved in the interaction (22). Substitutions of alanine for any of four arginine residues on the back face of UL42 resulted in substantial reductions in DNA binding without affecting the binding to peptide corresponding to the C terminus of Pol in vitro (31), while substitutions of lysine for arginine had little or no effect on DNA binding affinity (22). A UL42 mutant (Q282R) containing a substitution of arginine for a negatively charged glutamine residue on the back face of UL42 exhibited a fourfold increase in DNA binding without altering the interaction with the Pol C-terminal peptide in vitro (22). Therefore, the positively charged surface of UL42 is important for the interaction between UL42 and DNA. A question raised by these studies is whether UL42 could bind DNA so tightly as to affect HSV replication.Mutant viruses engineered to encode individual arginine-to-alanine substitution mutations in UL42 exhibit several phenotypes, including a delayed onset of viral DNA replication, reduced virus yields, and reduced fidelity of DNA replication (18). Recombinant viruses expressing UL42 with multiple substitutions of alanine for arginine residues exhibit even greater effects on viral DNA replication and virus yields (19). Thus, reducing DNA binding by UL42 deleteriously affects viral growth and DNA replication fidelity. However, these studies did not address whether increasing DNA binding by UL42 would have any effects on viral DNA replication, replication fidelity, or virus production.In this study we engineered two new UL42 mutant proteins (with the D270A/D271A or Q282R/D270A/D271A mutations) that contain less negative charge on the back face and examined the effects of these substitutions on DNA and Pol peptide binding. In addition, recombinant viruses were constructed to examine the effect of these multiple substitutions and the single Q282R substitution on virus production, DNA replication, and the fidelity of DNA replication.  相似文献   

12.
The assimilation and respiration of glucose by attached and free-living Pseudomonas fluorescens were compared. The attachment surfaces were polyvinylidene fluoride, polyethylene, and glass. Specific uptake of [14C]glucose was determined after bacterial biomass was measured by (i) microscopic counts or (ii) prelabeling of cells by providing [3H]leucine as substrate, followed by dual-labeling scintillation counting. The glucose concentration was 1.4, 3.5, 5.5, 7.6, or 9.7 μM. Glucose assimilation by cells which became detached from the surfaces during incubation with glucose was also measured after the detached cells were collected by filtration. The composition of the substratum had no effect on the amount of glucose assimilated by attached cells. Glucose assimilation by attached cells exceeded that by free-living cells by a factor of between 2 and 5 or more, and respiration of glucose by surface-associated cells was greater than that by free-living bacteria. Glucose assimilation by detached cells was greater than that by attached bacteria. Measurements of biomass by microscopic counts gave more consistent results that those obtained with dual-labeling, but in general, results obtained by both methods were corroborative.  相似文献   

13.
Cis-regulatory elements play important roles in tissue-specific gene expression and in the evolution of various phenotypes, and mutations in promoters and enhancers may be responsible for adaptations of species to environments. TRIM72 is a highly conserved protein that is involved in energy metabolism. Its expression in the heart varies considerably in primates, with high levels of expression in Old World monkeys and near absence in hominids. Here, we combine phylogenetic hypothesis testing and experimentation to demonstrate that mutations in promoter are responsible for the differences among primate species in the heart-specific expression of TRIM72. Maximum likelihood estimates of lineage-specific substitution rates under local-clock models show that relative to the evolutionary rate of introns, the rate of promoter was accelerated by 78% in the common ancestor of Old World monkeys, suggesting a role for positive selection in the evolution of the TRIM72 promoter, possibly driven by selective pressure due to changes in cardiac physiology after species divergence. We demonstrate that mutations in the TRIM72 promoter account for the differential myocardial TRIM72 expression of the human and the rhesus macaque. Furthermore, changes in TRIM72 expression alter the expression of genes involved in oxidative phosphorylation, which in turn affects mitochondrial respiration and cardiac energy capacity. On a broader timescale, phylogenetic regression analyses of data from 29 mammalian species show that mammals with high cardiac expression of TRIM72 have high heart rate, suggesting that the expression changes of TRIM72 may be related to differences in the heart physiology of those species.  相似文献   

14.
Small RNAs play essential regulatory roles in genome stability, development, and responses to biotic and abiotic stresses in most eukaryotes. In plants, the RNaseIII enzyme DICER-LIKE1 (DCL1) produces miRNAs, whereas DCL2, DCL3, and DCL4 produce various size classes of siRNAs. Plants also encode RNASE THREE-LIKE (RTL) enzymes that lack DCL-specific domains and whose function is largely unknown. We found that virus infection induces RTL1 expression, suggesting that this enzyme could play a role in plant–virus interaction. To first investigate the biochemical activity of RTL1 independent of virus infection, small RNAs were sequenced from transgenic plants constitutively expressing RTL1. These plants lacked almost all DCL2-, DCL3-, and DCL4-dependent small RNAs, indicating that RTL1 is a general suppressor of plant siRNA pathways. In vivo and in vitro assays revealed that RTL1 prevents siRNA production by cleaving dsRNA prior to DCL2-, DCL3-, and DCL4-processing. The substrate of RTL1 cleavage is likely long-perfect (or near-perfect) dsRNA, consistent with the RTL1-insensitivity of miRNAs, which derive from DCL1-processing of short-imperfect dsRNA. Virus infection induces RTL1 mRNA accumulation, but viral proteins that suppress RNA silencing inhibit RTL1 activity, suggesting that RTL1 has evolved as an inducible antiviral defense that could target dsRNA intermediates of viral replication, but that a broad range of viruses counteract RTL1 using the same protein toolbox used to inhibit antiviral RNA silencing. Together, these results reveal yet another level of complexity in the evolutionary battle between viruses and plant defenses.  相似文献   

15.
Three known genes guide circumferential migrations of pioneer axons and mesodermal cells on the nematode body wall. unc-5 affects dorsal migrations, unc-40 primarily affects ventral migrations, and unc-6 affects migrations in both directions. Circumferential movements still occur, but are misdirected whereas longitudinal movements are normal in these mutants. Pioneer growth cones migrating directly on the epidermis are affected; growth cones migrating along established axon fascicles are normal. Thus these genes affect cell guidance and not cell motility per se. We propose that two opposite, adhesive gradients guide circumferential migrations on the epidermis. unc-5, unc-6, and unc-40 may encode these adhesion molecules or their cellular receptors. Neurons have access to the basal lamina and the basolateral surfaces of the epidermis, but mesodermal cells contact only the basal lamina. These genes probably identify molecular cues on the basal lamina that guide mesodermal migrations. The same basal lamina cues, or perhaps related molecules on the epidermal cell surfaces, guide pioneer neurons.  相似文献   

16.
17.
The effects of insulin stimulation and muscle contractions on the subcellular distribution of GLUT4 in skeletal muscle have been studied on a preparation of single whole fibers from the rat soleus. The fibers were labeled for GLUT4 by a preembedding technique and observed as whole mounts by immunofluorescence microscopy, or after sectioning, by immunogold electron microscopy. The advantage of this preparation for cells of the size of muscle fibers is that it provides global views of the staining from one end of a fiber to the other and from one side to the other through the core of the fiber. In addition, the labeling efficiency is much higher than can be obtained with ultracryosections. In nonstimulated fibers, GLUT4 is excluded from the plasma membrane and T tubules. It is distributed throughout the muscle fibers with ~23% associated with large structures including multivesicular endosomes located in the TGN region, and 77% with small tubulovesicular structures. The two stimuli cause translocation of GLUT4 to both plasma membrane and T tubules. Quantitation of the immunogold electron microscopy shows that the effects of insulin and contraction are additive and that each stimulus recruits GLUT4 from both large and small depots. Immunofluorescence double labeling for GLUT4 and transferrin receptor (TfR) shows that the small depots can be further subdivided into TfR-positive and TfR-negative elements. Interestingly, we observe that colocalization of TfR and GLUT4 is increased by insulin and decreased by contractions. These results, supported by subcellular fractionation experiments, suggest that TfR-positive depots are only recruited by contractions. We do not find evidence for stimulation-induced unmasking of resident surface membrane GLUT4 transporters or for dilation of the T tubule system (Wang, W., P.A. Hansen, B.A. Marshall, J.O. Holloszy, and M. Mueckler. 1996. J. Cell Biol. 135:415–430).  相似文献   

18.
Growth and differentiation factor 5 (GDF5) plays a central role in bone and cartilage development by regulating the proliferation and differentiation of chondrogenic tissue. GDF5 is synthesized as a preproprotein. The biological function of the proregion comprising 354 residues is undefined. We identified two families with a heterozygosity for the novel missense mutations p.T201P or p.L263P located in the proregion of GDF5. The patients presented with dominant brachydactyly type C characterized by the shortening of skeletal elements in the distal extremities. Both mutations gave rise to decreased biological activity in in vitro analyses. The variants reduced the GDF5-induced activation of SMAD signaling by the GDF5 receptors BMPR1A and BMPR1B. Ectopic expression in micromass cultures yielded relatively low protein levels of the variants and showed diminished chondrogenic activity as compared to wild-type GDF5. Interestingly, stimulation of micromass cells with recombinant human proGDF5T201P and proGDF5L263P revealed their reduced chondrogenic potential compared to the wild-type protein. Limited proteolysis of the mutant recombinant proproteins resulted in a fragment pattern profoundly different from wild-type proGDF5. Modeling of a part of the GDF5 proregion into the known three-dimensional structure of TGFβ1 latency-associated peptide revealed that the homologous positions of both mutations are conserved regions that may be important for the folding of the mature protein or the assembly of dimeric protein complexes. We hypothesize that the missense mutations p.T201P and p.L263P interfere with the protein structure and thereby reduce the amount of fully processed, biologically active GDF5, finally causing the clinical loss of function phenotype.  相似文献   

19.
Bruck syndrome is a rare autosomal recessive connective tissue disorder characterized by fragile bones, joint contractures, scoliosis, and osteoporosis. The telopeptides of bone collagen I are underhydroxylated in these patients, leading to abnormal collagen cross-linking. Three point mutations in lysyl hydroxylase (LH) 2, the enzyme responsible for the hydroxylation of collagen telopeptides, have been identified in Bruck syndrome. As none of them affects the residues known to be critical for LH activity, we studied their consequences at the molecular level by analyzing the folding and catalytic properties of the corresponding mutant recombinant polypeptides. Folding and oligomerization of the R594H and G597V mutants were abnormal, and their activity was reduced by >95% relative to the wild type. The T604I mutation did not affect the folding properties, although the mutant retained only ∼8% activity under standard assay conditions. As the reduced activity was caused by a 10-fold increase in the Km for 2-oxoglutarate, the mutation interferes with binding of this cosubstrate. In the presence of a saturating 2-oxoglutarate concentration, the activity of the T604I mutant was ∼30% of that of the wild type. However, the T604I mutant did not generate detectable amounts of hydroxylysine in the N-terminal telopeptide of a recombinant procollagen I chain when coexpressed in insect cells. The low activity of the mutant LH2 polypeptides is in accordance with the markedly reduced extent of collagen telopeptide hydroxylation in Bruck syndrome, with consequent changes in the cross-linking of collagen fibrils and severe abnormalities in the skeletal structures.  相似文献   

20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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