首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
By combining Ce(IV)/EDTA with two pseudo-complementary peptide nucleic acids (pcPNAs), both strands in double-stranded DNA were site-selectively hydrolyzed at the target site. Either plasmid DNA (4361 bp) or its linearized form was used as the substrate. When two pcPNAs invaded into the double-stranded DNA, only the designated portion in each of the two strands was free from Watson–Crick base pairing with the counterpart DNA or the pcPNA. Upon the treatment of this invasion complex with Ce(IV)/EDTA at 37°C and pH 7.0, both of these single-stranded portions were selectively hydrolyzed at the designated site, resulting in the site-selective two-strand scission of the double-stranded DNA. Furthermore, the hydrolytic scission products were successfully connected with foreign double-stranded DNA by using ligase. The potential of these artificial systems for manipulation of huge DNA has been indicated.  相似文献   

2.
By combining Ce(IV)/EDTA with two pseudo-complementary peptide nucleic acids (pcPNAs), both strands in double-stranded DNA were site-selectively hydrolyzed at the target site. Either plasmid DNA (4361 bp) or its linearized form was used as the substrate. When two pcPNAs invaded into the double-stranded DNA, only the designated portion in each of the two strands was free from Watson-Crick base pairing with the counterpart DNA or the pcPNA. Upon the treatment of this invasion complex with Ce(IV)/EDTA at 37 degrees C and pH 7.0, both of these single-stranded portions were selectively hydrolyzed at the designated site, resulting in the site-selective two-strand scission of the double-stranded DNA. Furthermore, the hydrolytic scission products were successfully connected with foreign double-stranded DNA by using ligase. The potential of these artificial systems for manipulation of huge DNA has been indicated.  相似文献   

3.
The artificial restriction DNA cutter (ARCUT) method to cut double-stranded DNA at designated sites has been developed. The strategy at the base of this approach, which does not rely on restriction enzymes, is comprised of two stages: (i) two strands of pseudo-complementary peptide nucleic acid (pcPNA) anneal with DNA to form 'hot spots' for scission, and (ii) the Ce(IV)/EDTA complex acts as catalytic molecular scissors. The scission fragments, obtained by hydrolyzing target phosphodiester linkages, can be connected with foreign DNA using DNA ligase. The location of the scission site and the site-specificity are almost freely tunable, and there is no limitation to the size of DNA substrate. This protocol, which does not include the synthesis of pcPNA strands, takes approximately 10 d to complete. The synthesis and purification of the pcPNA, which are covered by a related protocol by the same authors, takes an additional 7 d, but pcPNA can also be ordered from custom synthesis companies if necessary.  相似文献   

4.
A new strategy for site-selective DNA hydrolysis, which takes advantage of the difference in reactivity between the phosphodiester linkages at the target site and the others, is presented. As the molecular scissors, homogeneous Ce(IV)/ethylenediamine-N,N,N′,N′-tetraacetate (EDTA) complex is used without being bound to any sequence-recognizing moiety. When a gap structure is formed at the target site by using two short oligonucleotides and the composite is treated with the Ce(IV)/EDTA complex at pH 7.0 and 37°C, the gap site in the substrate DNA is preferentially hydrolyzed over the double-stranded portion of the DNA. Site-selective DNA scission is also achieved by forming a bulge structure at the target site with the use of the appropriate oligonucleotide. These site-selective scissions are based on the following two factors: (i) the phosphodiester linkages in a single-stranded DNA are far more susceptible to the hydrolysis by the Ce(IV) complex than are the linkages in double-stranded DNA, and (ii) the phosphodiester linkages in the bulge sites are still more reactive than those in single-stranded DNA. In both cases, the addition of spermine significantly accelerates the scission.  相似文献   

5.
DNA looping is one of the key factors allowing proteins bound to different DNA sites to signal one another via direct contacts. We demonstrate that DNA looping can be generated in an arbitrary chosen site by sequence-directed targeting of double-stranded DNA with pseudocomplementary peptide-nucleic acids (pcPNAs). We designed pcPNAs to mask the DNA from cleavage by type IIs restriction enzyme PleI while not preventing the enzyme from binding to its primary DNA recognition site. Direct interaction between two protein molecules (one bound to the original recognition site and the other to a sequence-degenerated site) results in a totally new activity of PleI: it produces a nick near the degenerate site. The PNA-induced nicking efficiency varies with the distance between the two protein-binding sites in a phase with the DNA helical periodicity. Our findings imply a general approach for the fine-tuning of proteins bound to DNA sites well separated along the DNA chain.  相似文献   

6.
A new generation of PNAs, so-called pseudocomplementary PNAs (pcPNAs), which are able to target the designated sites on duplex DNA with mixed sequence of purines and pyrimidines via double-duplex invasion mode, has recently been introduced. It has been demonstrated that appropriate pairs of decameric pcPNAs block an access of RNA polymerase to the corresponding promoter. Here, we show that this type of PNAs protects selected DNA sites containing all four nucleobases from the action of restriction enzymes and DNA methyltransferases. We have found that pcPNAs as short as octamers form stable and sequence-specific complexes with duplex DNA in a very salt-dependent manner. In accord with a strand-invasion mode of complex formation, the pcPNA binding proceeds much faster with supercoiled than with linear plasmids. The double-duplex invasion complexes selectively shield specific DNA sites from BclI restriction endonuclease and dam methylase. The pcPNA-assisted protection against enzymatic methylation is more efficient when the PNA-binding site embodies the methylase-recognition site rather than overlaps it. We conclude that pcPNAs may provide the robust tools allowing to sequence-specifically manipulate DNA duplexes in a virtually sequence-unrestricted manner.  相似文献   

7.
Better understanding of proteins'' structure/function relationship and dissecting their functional domains are still challenges yet to be mastered. Site-directed mutagenesis approaches that can alter bases at precise positions on the gene sequence can help to reach this goal. This article describes an efficient strategy that can be applied not only for both deletion and substitution of target amino acids, but also for insertion of point mutations in promoter regions to study cis-regulating elements. This method takes advantage of the plasticity of the genetic code and the use of compatible restriction sites.Key words: site-directed mutagenesis, restriction site, cloning, PCRUnderstanding the proteins structure/function relationship and dissecting their functional domains is one of the biggest challenges to current proteomic studies.1 This is mainly achieved by site-directed mutagenesis experiments that can alter bases at precise positions on the gene sequence.2 Modifying DNA sequences has become feasible with PCR amplification.3 During the last decade, several strategies have been developed to simplify this approach and increase its efficiency.4 The introduction of a site-directed mutation can be realized by one or more PCR reactions. Most of the strategies used in site-directed mutagenesis are based on a substitution of a single base, which leads to a change in one amino acid. This article describes an efficient strategy that can be applied for either deletion or substitution of target amino acids. This strategy is based on performing PCR reactions to create a new restriction site in the sequence of origin, corresponding to the desired mutation. The choice of the restriction site to be created depends on the nature of the amino acid that one desires to introduce in the protein sequence. Since such restriction sites may extend beyond the mutated codon. The preservation of the other codon is done by taking advantage of the plasticity of the genetic code where one amino acid can be encoded by multiple codons.This method was performed in two steps (Fig. 1). In the first step, the DNA sequence of interest, cloned in a plasmid, served as a template for two PCR reactions. Two PCR products are generated. The first one consists of the beginning of the sequence, from the start codon to the mutagenized amino acid codon, where the forward primer bears the start codon region and the reverse primer bears the newly introduced restriction site at the same location of the mutagenized codon. The second PCR product consists of the end of the coding sequence, from the mutagenized amino acid codon to the stop codon. This fragment is generated using a forward primer bearing the same new restriction site as the first PCR product''s reverse primer, and a reverse primer bearing the stop codon region. The two PCR products were cloned separately into a vector in the appropriate orientation. In the second step, the cloning vector bearing the first PCR product was digested with a restriction enzyme site in the vector, and by the restriction enzyme corresponding to the restriction site created by the reverse primer used in the PCR reaction. The resulting fragment was cloned into the vector containing the second PCR fragment, predigested with same two restriction enzymes. The whole mutagenized coding sequence is reassembled by in-frame subcloning of the 3′ end of the coding sequence downstream the 5′ end. All the PCR products were generated using the high fidelity Pfu DNA Polymerase (Promega, Madison, WI USA). For any site-directed mutagenesis experiment, this two-step cloning procedure requires the use of appropriate PCR primers that harbor the desired mutation of the target amino acid. These primers are partially overlapping and contain a common or complementary restriction site enabling the in-frame assembly of the whole coding sequence.Open in a separate windowFigure 1Mutagenesis strategy by restriction enzyme site insertion. (A) In the first step, two PCR products were generated using the full length coding sequence as template. The mutation is carried by the two primers b and c, which are flanked by the same or compatible restriction enzyme''s site (white segment). Both PCR products are separately cloned in the cloning vector in the appropriate orientation. In the second step, the whole mutagenized coding sequence is reassembled by in-frame sub cloning of the 3′ end of the coding sequence downstream the 5′ end. (B) Substitution of threonine by arginine as a result of the insertion of a BglII restriction site. DNA sequencing is carried out to make sure that only the desired change is introduced in the coding sequence. (B-1) The sequence of the native cDNA. (B-2) the sequence of the mutagenized cDNA included BglII restriction site sequence.This approach has been used in a recent study to address the structure/function relationship of the STAS domain of the Arabidopsis thaliana Sultr1;2 sulfate transporter.5 A good example of this approach is the replacement of the threonine-serine couple at position 587–588 with an arginine-serine couple. The codon for threonine is: TGT, and that for arginine is: TCT. Serine can be encoded by both TCA and AGA codons. The chosen restriction site used for the reassembly of the whole coding sequence is that of the BglII enzyme: TCT AGA. The insertion of this restriction site enables the substitution of the Thr in position 587 with an Arg while preserving the serine residue in position 588. The BglII restriction site is introduced in the reverse primer and the forward primer used to generate the first and second PCR products respectively. The DNA sequence of the reassembled mutagenized cDNA was checked by sequencing. Than it was expressed, under pGAL1O promoter bearing by pYES2 vector, in yeast mutant deficient in sulfate transporter and the mutagenic protein was detected by imunodetection.Bioinformatic study reveals that this method can be applied to checked a large number of substitutions, insertions or deletions and that finding the right restriction site is not a limiting factor (data no shown).In conclusion, this article describes an efficient two-step procedure for site-directed mutagenesis using primers bearing a restriction site, which is absent from the sequence of origin. The primers flanked by sequences introducing the same or compatible restriction sites mediate the incorporation of the mutation at the selection site. The choice of the restriction site depends on the nature of the desired mutation: insertion, substitution or deletion of an amino acid in a particular position. This strategy can be also used to insert point mutations in promoter regions to study cis-regulating elements.  相似文献   

8.
A chemistry-based artificial restriction DNA cutter (ARCUT) was recently prepared from Ce(IV)/EDTA complex and a pair of pseudo-complementary peptide nucleic acids. This cutter has freely tunable scission-site and site specificity. In this article, homologous recombination (HR) in human cells was promoted by cutting a substrate DNA with ARCUT, and the efficiency of this bioprocess was optimized by various chemical and biological approaches. Of two kinds of terminal structure formed by ARCUT, 3'-overhang termini provided by 1.7-fold higher efficiency than 5'-overhang termini. A longer homology length (e.g. 698 bp) was about 2-fold more favorable than shorter one (e.g. 100 bp). When the cell cycle was synchronized to G2/M phase with nocodazole, the HR was promoted by about 2-fold. Repression of the NHEJ-relevant proteins Ku70 and Ku80 by siRNA increased the efficiency by 2- to 3-fold. It was indicated that appropriate combination of all these chemical and biological approaches should be very effective to promote ARCUT-mediated HR in human cells.  相似文献   

9.
A targeting method to insert genes at a previously characterized genetic locus to make plant transformation and transgene expression predictable is highly desirable for plant biotechnology. We report the successful targeting of transgenes to predefined soybean (Glycine max) genome sites using the yeast FLP-FRT recombination system. First, a target DNA containing a pair of incompatible FRT sites flanking a selection gene was introduced in soybean by standard biolistic transformation. Transgenic events containing a single copy of the target were retransformed with a donor DNA, which contained the same pair of FRT sites flanking a different selection gene, and a FLP expression DNA. Precise DNA cassette exchange was achieved between the target and donor DNA via recombinase-mediated cassette exchange, so that the donor DNA was introduced at the locus previously occupied by the target DNA. The introduced donor genes expressed normally and segregated according to Mendelian laws.Plant transformation has challenges such as random integration, multiple transgene copies, and unpredictable expression. Homologous recombination (Iida and Terada, 2005; Wright et al., 2005) and DNA recombinase-mediated site-specific integration (SSI) are promising technologies to address the challenges for placing a single copy of transgenes into a precharacterized site in a plant genome.Several site-specific DNA recombination systems, such as the bacteriophage Cre-lox and the yeast FLP-FRT and R-RS, have been used in SSI studies (Ow, 2002; Groth and Calos, 2003). A common feature of these systems is that each system consists of a recombinase Cre, FLP, or R and two identical or similar palindromic recognition sites, lox, FRT, or RS. Each recognition site contains a short asymmetric spacer sequence where DNA strand exchange takes place, flanked by inverted repeat sequences where the corresponding recombinase specifically binds. If two recognition sites are located in cis on a DNA molecule, the DNA segment can be excised if flanked by two directionally oriented sites or inverted if flanked by two oppositely oriented sites. If two recognition sites are located in trans on two different DNA molecules, a reciprocal translocation can happen between the two DNA molecules or the two molecules can integrate if at least one of them is a circular DNA (Ow, 2002; Groth and Calos, 2003).Single-site SSI can integrate a circular donor DNA containing one recognition site into a similar site previously placed in a plant genome. The integrated transgene now flanked by two recognition sites is vulnerable to excision. Transient Cre expression and the use of mutant lox sites to create two less compatible sites after integration helped reduce the subsequent excision in tobacco (Nicotiana tabacum; Albert et al., 1995; Day et al., 2000). A similar approach was used to produce SSI events in rice (Oryza sativa), and the transgene was proven stably expressed over generations (Srivastava and Ow, 2001; Srivastava et al., 2004; Chawla et al., 2006). Using a promoter trap to displace a cre gene in the genome with a selection gene from the donor, approximately 2% SSI was achieved in Arabidopsis (Arabidopsis thaliana; Vergunst et al., 1998).When two recognition sites located on a linear DNA molecule are similar enough to be recognized by the same recombinase but different enough to reduce or prevent DNA recombination from happening between them, the DNA segment between the two sites may not be easily excised or inverted. When a circular DNA molecule carrying the same two incompatible sites is introduced, the circular DNA can integrate by the corresponding recombinase at either site on the linear DNA to create a collinear DNA with four recognition sites, two from the original linear DNA and two from the circular DNA. DNA excision can subsequently occur between any pair of compatible sites to restore the two original DNA molecules or to exchange the intervening DNA segments between the two DNA molecules. This process, termed recombinase-mediated cassette exchange (RMCE), can be employed to integrate transgenes directionally into predefined genome sites (Trinh and Morrison, 2000; Baer and Bode, 2001).RMCE using two oppositely oriented identical RS sites, a donor containing the R recombinase gene and a third RS site to limit random integration, resulted in cassette exchange between the donor and a previously placed target in tobacco (Nanto et al., 2005). RMCE using both the Cre-lox and FLP-FRT systems improved RMCE frequency in animal cell cultures (Lauth et al., 2002). RMCE using two directly oriented incompatible FRT sites and transiently expressed FLP recombinase achieved cassette exchange between a target previously placed in the Drosophila genome and a donor introduced as a circular DNA (Horn and Handler, 2005). A gene conversion approach involving Cre-lox- and FLP-FRT-mediated SSI, RMCE, and homologous recombination was explored in maize (Zea mays; Djukanovic et al., 2006). RMCE using two oppositely oriented incompatible lox sites and transiently expressed Cre recombinase produced single-copy RMCE plants in Arabidopsis (Louwerse et al., 2007).To develop FLP-FRT-mediated RMCE in soybean (Glycine max), we created transgenic target lines containing a hygromycin resistance gene flanked by two directly oriented incompatible FRT sites via biolistic transformation. Single-copy target lines were selected and retransformed with a donor DNA containing a chlorsulfuron resistance gene flanked by the same pair of FRT sites. An FLP expression DNA was cobombarded to transiently provide FLP recombinase. RMCE events were obtained from multiple target lines and confirmed by extensive molecular characterization.  相似文献   

10.
By combining two strands of pseudo-complementary peptide nucleic acid (pcPNA) with S1 nuclease, a tool for site-selective and dual-strand scission of DNA/RNA hybrids has been developed. Both of the DNA and the RNA strands in the hybrids are hydrolyzed at desired sites to provide unique sticky ends. The scission fragments are directly ligated with other DNA/RNA hybrids by using T4 DNA ligase, resulting in the formation of desired recombinant DNA/RNA hybrids.  相似文献   

11.
An artificial restriction enzyme, which we developed recently by combining Ce(IV)/EDTA and peptide nucleic acids, was used for PCR-free construction of a fusion protein. The fusion protein was successfully expressed in mammalian cells. This artificial DNA cutter can be also applied to site-selective scission of huge DNAs. Promising features of this novel tool were concretely evidenced.  相似文献   

12.
Photo-responsive phosphoramidite monomers, which bear an azobenzene between acridine and the phosphoramidite unit, were synthesized, and incorporated into oligonucleotides. Upon UV irradiation, the azobenzene in the modified DNA efficiently isomerized from the trans isomer into the cis isomer. Although the Tm values of their duplexes with complementary DNA were not much changed by the isomerization, site-selective RNA scission was significantly accelerated by the UV irradiation when Mn(II) ion was used as the catalyst for RNA scission.  相似文献   

13.
Enzymes that modify DNA are faced with significant challenges in specificity for both substrate binding and catalysis. We describe how single hydrogen bonds between M.HhaI, a DNA cytosine methyltransferase, and its DNA substrate regulate the positioning of a peptide loop which is ∼28 Å away. Stopped-flow fluorescence measurements of a tryptophan inserted into the loop provide real-time observations of conformational rearrangements. These long-range interactions that correlate with substrate binding and critically, enzyme turnover, will have broad application to enzyme specificity and drug design for this medically relevant class of enzymes.Sequence-specific modification of DNA is essential for nearly all forms of life and contributes to a myriad of biological processes including gene regulation, mismatch repair, host defense, DNA replication, and genetic imprinting. Methylation of cytosine and adenine bases is a key epigenetic process whereby phenotypic changes are inherited without altering the DNA sequence (1). The central role of the bacterial and mammalian S-adenosylmethionine (AdoMet)2-dependent DNA methyltransferases in virulence regulation and tumorigenesis, respectively, have led these enzymes to be validated targets for antibiotic and cancer therapies (2, 3). However, AdoMet-dependent enzymes catalyze diverse reactions, and the design of potent and selective DNA methyltransferase inhibitors is particularly challenging (4, 5). The design of drugs that bind outside the active site is a particularly attractive means of inhibition for enzymes with common cofactors like AdoMet because off-target inhibition often leads to toxicity (6). Unfortunately, robust methods to identify and characterize such critical binding sites distal from the active site have not been developed.DNA methyltransferases bind to a particular DNA sequence, stabilize the target base into an extrahelical position within the enzyme active site, and transfer the methyl moiety from AdoMet to the DNA (7). During this process, dramatic changes in the DNA structure such as bending, base flipping, or the intercalation of residues into the recognition sequence are often accompanied by large scale protein rearrangements (8). Here we characterized a specific conformational rearrangement of M.HhaI, a model DNA cytosine C5 methyltransferase with a cognate recognition sequence of 5′-GCGC-3′. Many structures of M.HhaI are available at high resolution including an ensemble of complexes with either cognate or nonspecific DNA (9, 10). Reorganization of an essential catalytic loop (residues 80–100) is regulated by sequence-specific protein-DNA interactions that occur ∼28 Å away from the catalytic loop (Fig. 1). Our work quantifies the importance of such distal communication in sequence-specific DNA modification and provides plausible structural mechanisms.Open in a separate windowFIGURE 1.Loop interactions in M.HhaI. A, two superimposed structures of M.HhaI are shown with the catalytic loop highlighted. Enzymes are in blue (open conformer) and red (closed conformer) with the cofactor in orange, the flipped cytosine in green, and position 1 of the DNA recognition sequence colored by atom along with Cys-81, Ile-86, and Arg-240. The large blue arrow shows the long-range structural communication between Arg-240 and the catalytic loop. B, close-up of the interactions between Arg-240, Ile-86, and position 1 of the recognition sequence. The flipped cytosine is in green. Removal of N2 from the guanosine with an inosine base maintains near cognate loop motion while removal of O6 with a 2AP base has almost no loop motion.DNA-dependent positioning of the catalytic loop in M.HhaI was first observed crystallographically; cognate DNA stabilizes the loop-closed conformer, while nonspecific DNA leaves the loop in the open conformer (9, 10). Correct positioning of this loop is essential for catalysis because C81, the active site nucleophile that attacks the target cytosine base at the C6 position (supplemental Fig. S1), is ∼9.6 Å away in the loop-open conformer (Fig. 1A). Populating the closed conformer of the loop is essential for tight DNA binding and stabilizing the target cytosine that is flipped out of the DNA duplex (1113). Using stopped-flow fluorescence spectroscopy to monitor the environment of tryptophan (Trp) residues inserted into the catalytic loop, we recently observed reorganization of this loop upon DNA binding in the absence of cofactor using the M.HhaI mutants W41F, W41F/K91W, and W41F/E94W (12). Loop positioning and the interconversion between the open and closed conformers, as determined from the intensities and rates of change in fluorescence signal are highly dependent on DNA sequence and confirm that cognate DNA stabilizes the loop-closed conformer whereas nonspecific DNA stabilizes the open conformer.In this study, W41F/K91W and W41F/E94W M.HhaI were preincubated with cognate (COG), non-cognate (NC), or nonspecific (NS) DNA and mixed with cofactor or cofactor product, AdoMet and S-adenosylhomocysteine (AdoHcy), respectively, in a stopped-flow apparatus. Differences in observed fluorescence intensity are indicative of shifts in the populations of the various loop conformers; no observable fluorescence change suggests no significant change in population and thus, essentially no loop positioning to the closed conformer. Non-cognate and cognate sequences are nearly identical but differ by a single base change, whereas the nonspecific DNA substrate has no similarity to the cognate sequence. As a methyltransferase searches for the cognate site within a genome, it must encounter both non-cognate and nonspecific DNA sequences and be able to distinguish these from the cognate sequence. We examine both binding, using the cofactor product AdoHcy, and catalysis, using the AdoMet cofactor of M.HhaI, with these various DNA substrates.  相似文献   

14.
The five EcoRI2 restriction sites in bacteriophage lambda DNA have been mapped at 0.445, 0.543, 0.656, 0.810, and 0.931 fractional lengths from the left end of the DNA molecule. These positions were determined electron-microscopically by single-site cleavage of hydrogen-bonded circular λ DNA molecules and by cleavage of various DNA heteroduplexes between λ DNA and DNA from well defined λ mutants. The DNA lengths of the EcoRI fragments are in agreement with their electrophoretic mobility on agarose gels but are not in agreement with their mobilities on polyacrylamide gels. These positions are different from those previously published by Allet et al. (1973). Partial cleavage of pure λ DNA by addition of small amounts of EcoRI endonuclease does not lead to random cleavage between molecules. Also, the first site cleaved is not randomly distributed among the five sites within a molecule. The site nearest the right end is cleaved first about ten times more frequently than either of the two center sites.  相似文献   

15.
Photo-responsive phosphoramidite monomers, which bear an azobenzene between acridine and the phosphoramidite unit, were synthesized, and incorporated into oligonucleotides. Upon UV irradiation, the azobenzene in the modified DNA efficiently isomerized from the trans isomer into the cis isomer. Although the T(m) values of their duplexes with complementary DNA were not much changed by the isomerization, site-selective RNA scission was significantly accelerated by the UV irradiation when Mn(II) ion was used as the catalyst for RNA scission.  相似文献   

16.
Summary The activity of the EcoK DNA restriction system of Escherichia coli reduces both the plating efficiency of unmodified phage and the transforming ability of unmodified pBR322 plasmid DNA. However, restriction can be alleviated in wild-type cells, by UV irradiation and expression of the SOS response, so that 103-to 104-fold increases in phage growth and fourfold increases in plasmid transformation occurred with unmodified DNA. Restriction alleviation was found to be a transient effect because induced cells, which initially failed to restrict unmodified plasmid DNA, later restricted unmodified phage . Although the SOS response was needed for restriction alleviation, constitutive SOS induction, elicited genetically with a recA730 mutation, did not alleviate restriction and UV irradiation was still needed. A hitherto unsuspected involvement of the umuDC operon in this alleviation of restriction is characterized and, by differential complementation, was separated from the better known role of umuDC in mutagenic DNA repair. The need for cleavage of UmuD for restriction alleviation was shown with plasmids encoding cleavable, cleaved, and non-cleavable forms of UmuD. However, UV irradiation was still needed even when cleaved UmuD was provided. The possibility that restriction alleviation occurs by a general inhibition of the EcoK restriction/modification complex was tested and discounted because modification of was not reduced by UV irradiation. An alternative idea, that restriction activity was competitively reduced by an increase in EcoK modification, was also discounted by the lack of any increase in the modification of Ral, a naturally undermodified phage. Other possible mechanisms for restriction alleviation are discussed.  相似文献   

17.
Methods for the extraction of PCR-quality DNA from environmental soil samples by using pairs of commercially available kits were evaluated. Coxiella burnetii DNA was detected in spiked soil samples at <1,000 genome equivalents per gram of soil and in 12 (16.4%) of 73 environmental soil samples.The detection of pathogenic organisms in the environment often relies on PCR analysis of DNA purified from environmental soil (6). For effective detection, a reliable method to obtain PCR-quality DNA from soil is necessary. Although a variety of complex techniques have been effective for specific soil samples (1-3, 7, 8), it is not clear which methods would be the best for the wide variety of samples encountered in a large-scale environmental sampling study. In addition, many published techniques would be difficult to use on a large number of samples (1-3, 7, 8).This study evaluates the abilities of commercially available DNA extraction kits to provide DNA from environmental soil samples that are suitable for PCR detection of Coxiella burnetii. C. burnetii is an obligate intracellular, Gram-negative, zoonotic pathogen and the causative agent of Q fever (5). It is classified as a category B agent of bioterrorism by the CDC.Three commercially available DNA purification kits were evaluated. Twenty different soil samples obtained from diverse locations in the southeastern United States were used for testing. These samples consisted of light sandy soil and were all initially processed through one of three DNA purification kits, the UltraClean soil DNA isolation kit (MoBio Laboratories, Carlsbad CA), the QIAamp DNA minikit (Qiagen, Valencia, CA), or the QIAamp DNA stool minikit (Qiagen), or through a combination of two of the kits used sequentially. Thus, all 20 samples were each processed through nine extraction protocols. To process soil samples, five grams of soil was mixed with 10 to 30 ml of phosphate-buffered saline (PBS) to create a homogenized slurry. Samples were mixed for 1 h at room temperature and then centrifuged for 5 min at 123 × g. The supernatant was removed and centrifuged at 20,000 × g for 15 min. The supernatant was then carefully discarded and the pellet resuspended in 1 ml of PBS.For the UltraClean soil kit, 700 μl of the resuspended soil extraction pellet was processed by the manufacturer''s alternative protocol (for maximum yields). For preps done using the QIAamp DNA minikit (tissue protocol) and the QIAamp stool kit (stool protocol), 700 μl (high volume) of the soil extract was processed according to the instructions for the particular kit. For 17 of the samples the tissue protocol and stool protocol were applied using only 200 μl of the soil extract (low volume). For all of the kits, the final elutions were performed with 55 μl of water.To further purify the products of the commercial DNA isolation kits, eluates were passed through a second round of extraction. When the MoBio UltraClean kit was used for the second round of extraction, eluates were added to the bead-containing tubes and mixed with 60 μl of solution 1 and 200 μl of the MoBio inhibitor removal solution (IRS). The manufacturer''s protocol was then followed. When the QIAamp tissue protocol was utilized for the second round of extraction, eluates were diluted to 200 μl with water and then mixed with 200 μl of buffer ATL plus 200 μl of buffer AL and then incubated at 70°C for 10 min. Following this step, the manufacturer''s protocol was followed. When the QIAamp stool protocol was used for the second round of extraction, eluates were mixed with 1.2 ml of the ASL buffer, followed by addition of the InhibitEX tablet. The manufacturer''s protocol was then followed.PCR inhibition in all of the DNA samples was then evaluated by running a quantitative PCR that detects the IS1111 gene from C. burnetii (4). PCRs were run on 200 genome equivalents of C. burnetii (strain Nine Mile Phase 1) DNA. Reaction mixtures spiked with 1-μl aliquots of the environmental DNA samples were compared to reaction mixtures spiked with 1 μl of water. Inhibition was considered present if the DNA sample caused an increase of 1 in the threshold cycle value.Use of the MoBio UltraClean procedure by itself resulted in removal of inhibitors from 35% of the samples, whereas after use of the Qiagen tissue protocol (high volume) only 4% of the samples were free of inhibition (Fig. (Fig.1).1). The Qiagen stool kit (high volume) resulted in 96% of the samples showing lack of inhibition with a low volume of soil eluate and 62.5% of the samples when the high volume was used. The DNA extracted from these three kits was then used as starting material for a subsequent DNA extraction step using the same set of three commercial kits. The MoBio UltraClean kit followed by the Qiagen stool kit eliminated inhibition in all samples, as did these two kits when used in the reverse order, even if the Qiagen stool kit was loaded with 700 μl of material (high volume). When a low volume of starting material was used, combinations of the two Qiagen kits also removed inhibitors from 100% of the samples when either the Qiagen tissue protocol was used first or the Qiagen stool protocol was used first (Fig. (Fig.1).1). The raw data for all of the inhibition assays are included as supplemental data (see Table S1 in the supplemental material).Open in a separate windowFIG. 1.Twenty environmental soil samples were used for the isolation of DNA with the indicated protocols. The samples were then tested for the ability to inhibit an IS1111 PCR with C. burnetii Nine Mile DNA as template. The percentages of samples that did not show any inhibition are indicated.To determine the yield of DNA obtained by the various protocols, nine aliquots (5 g each) of a single rich organic soil sample were each mixed with 5 ml PBS, spiked with 1 × 106 Nine Mile Phase 2 C. burnetii organisms, and then processed by the nine (high-volume) extraction protocols described above. An additional 1 × 106 Nine Mile Phase 2 C. burnetii organisms were used directly in the Qiagen tissue protocol to prepare DNA for the purpose of determining the exact amount of C. burnetii input into the assays. The quantitative IS1111 PCR assay (4) was used to determine the yield of C. burnetii DNA by using the various methods for processing soil. The yield was calculated by dividing the number of genome equivalents of C. burnetii DNA obtained from the spiked soil samples by the number of genome equivalents obtained when C. burnetii was included directly in the Qiagen tissue protocol. A common feature of all of the protocols was that they all produced a low yield of C. burnetii DNA when purified from a complex soil mixture (Fig. (Fig.2).2). The yields ranged from 0.02% to 4.3% and were variable. Although the 4.3% yield obtained when the stool kit was used alone was the highest on average, the high variability observed with these extractions suggests that most of these protocols provide similar yields. The stool kit followed by the MoBio kit clearly resulted in the lowest yield.Open in a separate windowFIG. 2.Five-gram aliquots of a single soil sample were all spiked with approximately 1 × 106 C. burnetii Phase 2 Nine Mile strain cells. The samples were then subjected to the indicated extraction protocol(s). The resulting DNA was tested for inhibition, and then the genome equivalents of C. burnetii DNA were determined by quantitative IS1111 PCR. The exact input amount of C. burnetii was determined by running an aliquot directly through the QIAamp tissue protocol followed by IS1111 PCR. Yield was calculated as genome equivalents obtained from the spiked soil samples divided by the genome equivalents obtained from the direct extraction through the QIAamp tissue protocol. Values represent the mean ± standard deviation of five experiments. Statistically significant differences (Student''s t test) were found between stool versus MoBio plus stool kits (P = 0.05), stool plus tissue versus MoBio plus stool kits (P = 0.01), and stool plus tissue versus tissue plus MoBio kits (P = 0.03). For the protocol using the stool kit followed by the MoBio kit the yield was significantly different from stool, stool plus tissue, MoBio plus tissue, and MoBio protocols (P < 0.05).Although these yields are low, the IS1111 PCR assay used to detect C. burnetii DNA amplifies a multicopy gene, and the assay can detect a single genome equivalent (4). This suggests that these protocols are adequate for the detection of C. burnetii in soil samples with 500 to 2,000 organisms per gram of soil. To test this, a 5-g sample of organic soil was spiked with 800 C. burnetii organisms per gram, and the DNA was extracted using the MoBio UltraClean kit followed by the QIAamp stool protocol. C. burnetii DNA was detected after 38 cycles using the IS1111 PCR assay.While these results are focused on soil samples, the procedures described also work well on vacuum samples and sponge wipe samples (data not shown). Based on removal of inhibitors and yield, our data suggest that the QIAamp tissue protocol (high volume) followed by the QIAamp stool protocol and the MoBio UltraClean kit followed by the QIAamp stool protocol are both suitable for extraction of DNA from environmental soil samples. To test the application of the latter method to a larger number of samples, 73 bulk soil samples from the southeastern United States were processed according to this method. Inhibition was removed from all 73 samples, and 12 of the samples were positive in the C. burnetii IS1111 PCR assay. This suggests that this practical method for extraction of PCR-quality DNA can be successfully used to detect DNA from C. burnetii and other pathogens in large numbers of environmental samples.   相似文献   

18.
The map of the seven sites for the restriction endonuclease HindIII3 and the single site for endo R.HpaII on PM2 DNA was determined. This map was oriented with respect to the denaturation map of this DNA (Brack et al., 1975) by partial denaturation mapping of the fragments. A new method for localizing restriction fragments by DNA-DNA hybridization and electron microscopy is described.  相似文献   

19.
An apparently full-length complementary DNA copy of in vitro polyadenylated MS2 RNA was synthesized with avian myeloblastosis virus RNA-dependent DNA polymerase. After the MS2 RNA template was removed from the complementary DNA strand with T1 and pancreatic RNase digestion, the complementary DNA became a good template for the synthesis of double-stranded MS2 DNA with Escherichia coli DNA polymerase I. We then constructed molecular chimeras by inserting the double-stranded MS2 DNA into the PstI restriction endonuclease cleavage site of the E. coli plasmid pBR322 by means of the poly(dA)· poly(dT) tailing procedure. An E. coli transformant carrying a plasmid with a nearly full-length MS2 DNA insertion, called pMS2-7, was chosen for further study. Correlation between the restriction cleavage site map of pMS2-7 DNA and the cleavage map predicted from the primary structure of MS2 RNA, and nucleotide sequence analysis of the 5′ and 3′ end regions of the MS2 DNA insertion, showed that the entire MS2 RNA had been faithfully copied, and that, except for 14 nucleotides corresponding to the 5′-terminal sequence of MS2 RNA, the fulllength DNA copy of the viral genetic information had been inserted into the plasmid. Restriction endonuclease analysis of the chimera plasmid DNA also revealed the presence of an extra DNA insertion which was identified as the translocatable element IS13 (see following paper).  相似文献   

20.
Sequence-specific DNA-binding molecules such as triple helix-forming oligonucleotides (TFOs) provide a means for inducing site-specific mutagenesis and recombination at chromosomal sites in mammalian cells. However, the utility of TFOs is limited by the requirement for homopurine stretches in the target duplex DNA. Here, we report the use of pseudo-complementary peptide nucleic acids (pcPNAs) for intracellular gene targeting at mixed sequence sites. Due to steric hindrance, pcPNAs are unable to form pcPNA–pcPNA duplexes but can bind to complementary DNA sequences by Watson–Crick pairing via double duplex-invasion complex formation. We show that psoralen-conjugated pcPNAs can deliver site-specific photoadducts and mediate targeted gene modification within both episomal and chromosomal DNA in mammalian cells without detectable off-target effects. Most of the induced psoralen-pcPNA mutations were single-base substitutions and deletions at the predicted pcPNA-binding sites. The pcPNA-directed mutagenesis was found to be dependent on PNA concentration and UVA dose and required matched pairs of pcPNAs. Neither of the individual pcPNAs alone had any effect nor did complementary PNA pairs of the same sequence. These results identify pcPNAs as new tools for site-specific gene modification in mammalian cells without purine sequence restriction, thereby providing a general strategy for designing gene targeting molecules.  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