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 46 毫秒
1.
A label-free electrochemical method for the detection of DNA-PNA hybridization using a water-soluble, ferrocene-functionalized polythiophene transducer and single-stranded PNA probes on the nanogold modified electrode is investigated. Nanogold modified electrodes can largely increase the immobilization amount of ss-PNA capture probe and lead to an increase of the electrical signal. The ferrocene-containing cationic polythiophene do not interact electrostatically with the PNA probes due to the absence of the anionic phosphate groups on the PNA probes. But after DNA-PNA hybridization, cationic polythiophene is adsorbed on the DNA backbone, giving a clear hybridization detection signal in differential pulse voltammetry (DPV). Very good discrimination against non-complementary DNA and four-base mismatch DNA is observed. These studies show that the proposed method can provide an alternative for expanding the range of detection methods available for DNA hybridization.  相似文献   

2.
Electrochemical impedance measurements were used for the detection of single-strand DNA sequences using a peptide nucleic acid (PNA) probe layer immobilized onto Si/SiO2 chips. An epoxysilane layer is first immobilized onto the Si/SiO2 surface. The immobilization procedure consists of an epoxide/amine coupling reaction between the amino group of the PNA linker and the epoxide group of the silane. A 20-nucleotide sequence of PNA was used. Impedance measurements allow for the detection of the changes in charge distribution at the oxide/solution interface following modifications to the oxide surface. Due to these modifications, there are significant shifts in the semiconductor's flat-band potential after immobilization and hybridization. The results obtained using this direct and rapid approach are supported by fluorescence measurements according to classical methods for the detection of nucleic acid sequences.  相似文献   

3.
Objectives: Peptide nucleic acid (PNA) probes hybridize to denatured telomeric sequences in cells permeabilized in hot formamide. In reported protocols, the hybridization was conducted in solutions with high formamide concentrations to avoid the DNA renaturation that can hamper binding of the oligo‐PNA probe to specific sequences. We postulated that telomeric DNA, confined in the nuclear microvolume, is not able to properly renature after hot formamide denaturation. Therefore, to improve hybridization conditions between the probe and the target sequences, it might be possible to add probe to sample after the complete removal of formamide. Materials and methods: After telomeric DNA denaturation in hot formamide solution and several washes to remove the ionic solvent, cells were hybridized overnight at room temperature with human telomere‐specific PNA probe conjugated with Cy5 fluorochrome, Cy5‐OO‐(CCCTAA)3. After stringency washes and staining with ethidium bromide, the cells were analysed by flow cytometry and by using a confocal microscope. Results: Using three continuous cell lines, different in DNA content and telomere length, and resting human peripheral blood T and B lymphocytes, we demonstrated that the oligo‐PNA probe hybridized to telomeric sequences after complete removal of formamide and that in the preserved nucleus, telomeric sequence denaturation is irreversible. Conclusion: According to our experience, oligo‐PNA binding results is efficient, specific and proportional to telomere length. These, our original findings, can form the technological basis of actual in situ hybridization on preserved whole cells.  相似文献   

4.
DNA sensors have a wide scope of applications in the present and emerging medical and scientific fields, such as medical diagnostics and forensic investigations. However, much research-to-date on DNA sensor development has focused on short target DNA strands as model genes. In this communication we study the effect of the length of oligonucleotide probe and target strands as a significant step towards real world applications for DNA detection. The sensor technology described uses the conducting polymer polypyrrole as both a sensing element and transducer of sensing events - namely the hybridization of complementary target oligonucleotide to probe oligonucleotide. Detection is performed using electrical impedance spectroscopy. Initially sensor development is performed, wherein we demonstrate an improvement in stability and sensitivity as well as show a reduction in non-specific DNA binding for fabricated sensors, through use of a specific dopant and post-growth treatment. Subsequently, we show that longer target DNA strands display increased response, as do sensors containing longer probe DNA strands. It is suggested that these results are a feature of the increase in negative charges associated with the longer DNA strands. The results of this comparative study are aimed to guide future design of analogous sensors.  相似文献   

5.
Mixed monolayers of the surface-active lipopeptide surfactin-C15 and various lipids differing by their chain length (DMPC, DPPC, DSPC) and polar headgroup (DPPC, DPPE, DPPS) were investigated by atomic force microscopy (AFM) in combination with molecular modeling (Hypermatrix procedure) and surface pressure-area isotherms. In the presence of surfactin, AFM topographic images showed phase separation for each surfactin-phospholipid system except for surfactin-DMPC, which was in good agreement with compression isotherms. On the basis of domain shape and line tension theory, we conclude that the miscibility between surfactin and phospholipids is higher for shorter chain lengths (DMPC > DPPC > DSPC) and that the polar headgroup of phospholipids influences the miscibility of surfactin in the order DPPC > DPPE > DPPS. Molecular modeling data show that mixing surfactin and DPPC has a destabilizing effect on DPPC monolayer while it has a stabilizing effect towards DPPE and DPPS molecular interactions. Our results provide valuable information on the activity mechanism of surfactin and may be useful for the design of surfactin delivery systems.  相似文献   

6.
The adsorption of a range of single chain zwitterionic phosphocholine surfactants (C(n)P(m)C) at the air/liquid interface has been studied by a combination of surface tension and neutron reflectivity. The critical micellar concentration (CMC) for C(n)PC (or C(n)P(2)C), where n varied from 12, 14 to 16, was found to be 0.91, 0.14, and 1.2 x 10(-2) mM respectively, and followed the same trend as observed for other zwitterionic and non-ionic surfactants. The area per molecule at the CMC, A(cmc), for C(n)PC was found to remain constant between 50 and 53 A(2), indicating that the increase in the alkyl chain length had little effect on A(cmc) at the interface. The neutron reflection measurement also showed an almost constant layer thickness (tau) of 20+/-2 A from all the alkyl chain deuterated PC surfactants (dC(n)hPC) in null reflecting water (NRW), suggesting that the alkyl chains of the surfactant responded to changes in either chain length or solution concentration by varying their angle of tilt. In contrast, increasing the length of head group linker between P and N atoms in C(12)P(m)C, where m=2, 4, to 6, resulted in a much slower decrease of CMC from 0.91, 0.7, to 0.5 mM, consistent with a different contribution to the free energy of micellization. A(cmc) for C(12)P(m)C did not vary when m was increased from 2 to 4, and this observation together with the thickness of the head group region indicated an almost perpendicular projection of the head group in C(12)P(2)C and C(12)P(4)C. A further increase in m to 6 resulted in an A(cmc) of 70 A(2). This increase in A(cmc) however did not result in any change in either the total layer thickness or the fraction of the head group region submerged in the aqueous subphase, suggesting that the head group in C(12)P(6)C was bent away from the surface normal direction. Both increase in temperature from 25 to 40 degrees C and the addition of 0.1 M NaCl had little effect on the area per molecule or the thickness of C(12)P(m)C surfactant layer, showing that the C(12)P(m)C series behaved like C(n)P(2)C series. The main conclusion from this study is that for all the C(n)P(m)C surfactants studied, change in m or n has little effect on the total thickness, the thickness of the alkyl chain or that of the head group region.  相似文献   

7.
Abstract A repeated DNA sequence was isolated from Fusobacterium necrophorum biotype AB, strain FnS1. The repeated sequence shared considerable homology with the transposase gene from the Pseudomonas syringiae insertion sequence IS801. The repeat sequence was used together with a 16S ribosomal RNA gene probe to type F. necrophorum isolates using restriction fragment length polymorphisms. The probes revealed differences between several clinical isolates and will be useful tools to study the epidemology of ovine foot abscess and other diseases caused by F. necrophorum .  相似文献   

8.
This paper explores the effects of structural modifications on the fast dynamics of DNA and the ability of time-resolved Stokes shift spectroscopy to measure those changes. The time-resolved Stokes shift of a synthetic coumarin base-pair replacement within an oligomer is measured between 40 ps and 40 ns. Comparisons are made between 17mers without modification, with a deleted base near the coumarin and with the coumarin placed near the end of the oligomer. The deletion of a next-to-nearest-neighbor base pair does not change the subnanosecond dynamics, but does cause an additional motion with a time constant of ~20 ns. A candidate for this motion is the flipping of the abasic sugar out of the helix and the concomitant intrusion of water into the interior of the helix. A nearby chain end causes little change in the dynamics after 1 ns but leads to a reduction in the amplitude of the dynamics between 40 ps and 1 ns. We suggest that at the chain end, where DNA on one side of the probe has been replaced by water, the charge- stabilizing dynamics have the same overall amplitude, but that much of the relaxation occurs before the start of the measurement time window.  相似文献   

9.
The major problem of using somatic mutations as markers of malignancy is that the clinical samples are frequently containing a trace amounts of mutant allele in a large excess of wild-type DNA. Most methods developed thus far for the purpose of tickling this difficult problem require multiple procedural steps that are laborious. We report herein the development of a rapid and simple protocol for detecting a trace amounts of mutant K-ras in a single tube, one-step format. In a capillary PCR, a 17mer peptide nucleic acid (PNA) complementary to the wild-type sequence and spanning codons 12 and 13 of the K-ras oncogene was used to clamp-PCR for wild-type, but not mutant alleles. The designated PNA was labeled with a fluorescent dye for use as a sensor probe, which differentiated all 12 possible mutations from the wild-type by a melting temperature (Tm) shift in a range of 9 to 16°C. An extension temperature of 60°C and an opposite primer 97 nt away from the PNA were required to obtain full suppression of wild-type PCR. After optimization, the reaction detected mutant templates in a ratio of 1:10000 wild-type alleles. Using this newly devised protocol, we have been able to detect 19 mutants in a group of 24 serum samples obtained from patients with pancreatic cancer. Taken together, our data suggest that this newly devised protocol can serve as an useful tool for cancer screening as well as in the detection of rare mutation in many diseases.  相似文献   

10.
As long-chain alcohol dehydrogenases are not easily available and seldom reported enzymes, it is worthwhile to appraise the potential of well known dehydrogenases, like horse liver alcohol dehydrogenases (HLAD), for the oxidation of long-chain aliphatic alcohols. Oxidation of docosanol (C22) and tetracosanol (C24) is of technological relevance within an industrial platform for the fractionation and upgrading of tall-oil from the Kraft pulping process. Results are presented on the characterization of free and immobilized HLAD with respect to their potential for oxidizing long-chain aliphatic alcohols. Enzyme activity with respect to chain length and pH is presented. Activity for both free and immobilized HLAD increased with pH up to 8.8, but behavior with respect to chain length varied from one biocatalyst to the other. Even though both biocatalysts were less active towards very long-chain aliphatic alcohols, immobilized HLAD had an activity on docosanol and tetracosanol higher than 50% of the value obtained with ethanol, butanol and octanol, which is encouraging and has not been previously reported. Investigation on thermophilic sources and further immobilization strategies are underway to obtain more active and stable catalysts amenable for working at high temperatures which is quite relevant in this case due to the poor solubility of substrates.  相似文献   

11.
The effect of serum on cell growth and monoclonal antibody (MAb) productivity was studied in a repeated fedbatch mode using both free-suspended and immobilized S3H5/gamma2bA2 hybridoma cells. In the suspension culture, serum influenced the cell growth rate but not the specific MAb productivity. The average specific growth rate of the suspension culture in medium containing 10% serum was approximately 0.99 +/- 0.12 day(-1) (+/-standard deviation), while that in medium containing 1% serum was approximately 0.73 +/- 0.12 day(-1). The specific MAb productivity was almost constant at 3.69 +/- 0.57 mug/10(6) cells/day irrespective of serum concentration reached a maximum at ca. 1.8 x 10(6) cells/mL of medium in 10% serum medium, and the cell concentration was gradually reduced to 1%. The specific MAb productivity of the immobilized cells was more than three times higher than that of the free-suspended cells. The amount of serum in the medium did not influence the specific MAb production rate of the immobilized cells. The maintenance of high cell concentration and the enhanced specific MAb productivity of the immobilized cell culture resulted in a higher volumetric MAb productivity. In addition, MAb yield in the immobilized cell culture with medium containing 1% serum was 2.2 mg/mL of serum, which was approximately three times higher than that in the suspension culture.  相似文献   

12.
The detergents, alkyltrimethylammonium bromide, N-alkyl-N, N-dimethyl-3-ammonio-1-propanesulfonate (zwittergent), alkane sulfonate, alkylsulfate, alkyl-beta-D-glucopyranoside, alkyl-beta-D-maltoside, dodecanoyl-N-methylglucamide, polyethylene glycol monoalkyl ether and Triton X-100, all produce a concentration-dependent acceleration of the slow passive transbilayer movement of NBD-labeled phosphatidylcholine in the human erythrocyte membrane. Above a threshold concentration, which was well below the CMC and characteristic for each detergent, the flip rate increases exponentially upon an increase of the detergent concentration in the medium. The detergent-induced flip correlates with reported membrane-expanding effects of the detergents at antihemolytic concentrations. From the dependence of the detergent concentration required for a defined flip acceleration on the estimated membrane volume, membrane/water partition coefficients for the detergents could be determined and effective detergent concentrations in the membrane calculated. The effective membrane concentrations are similar for most types of detergents but are 10-fold lower for octaethylene glycol monoalkyl ether and Triton X-100. The effectiveness of a given type of detergent is rather independent of its alkyl chain length. Since detergents do not reduce the high temperature dependence of the flip process the detergent-induced flip is proposed to be due to an enhanced probability of formation of transient hydrophobic structural defects in the membrane barrier which may result from perturbation of the interfacial region of the bilayer by inserted detergent molecules.  相似文献   

13.
Rachwal PA  Brown T  Fox KR 《Biochemistry》2007,46(11):3036-3044
G-Rich sequences are known to form four-stranded structures that are based on stacks of G-quartets, and sequences with the potential to adopt these structures are common in eukaryotic genomes. However, there are few rules for predicting the relative stability of folded complexes that are adopted by sequences with different-length G-tracts or variable-length linkers between them. We have used thermal melting, circular dichroism, and gel electrophoresis to examine the topology and stability of intramolecular G-quadruplexes that are formed by sequences of the type d(GnT)4 and d(GnT2)4 (n = 3-7) in the presence of varying concentrations of sodium and potassium. In the presence of potassium or sodium, d(GnT)4 sequences form intramolecular parallel complexes with the following order of stability: n = 3 > n = 7 > n = 6 > n = 5 > n = 4. d(G3T)4 is anomalously stable. In contrast, the stability of d(GnT2)4 increases with the length of the G-tract (n = 7 > n = 6 > n = 5 > n = 4 > n = 3). The CD spectra for d(GnT)4 in the presence of potassium exhibit positive peaks around 260 nm, consistent with the formation of parallel topologies. These peaks are retained in sodium-containing buffers, but when n = 4, 5, or 6, CD maxima are observed around 290 nm, suggesting that these sequences [especially d(G5T)4] have some antiparallel characteristics. d(G3T2)4 adopts a parallel conformation in the presence of both sodium and potassium, while all the other d(GnT2)4 complexes exhibit predominantly antiparallel features. The properties of these complexes are also affected by the rate of annealing, and faster rates favor parallel complexes.  相似文献   

14.
This paper demonstrates by means of FTIR/ATR analysis that water molecules intercalate at different extents in the acyl chain region of lipid membranes in correlation with the hydration of the phosphate groups.This correlation is sensible to the chain length, the presence of double bonds and the phase state of the lipid membrane.The presence of carbonyl groups CO modifies the profile of hydration of the two regions as observed from the comparison of DMPC and 14:0 Diether PC.The different water populations in lipid interphases would give arrangements with different free energy states that could drive the interaction of biological effectors with membranes.  相似文献   

15.
The ability to selectively target the harmful microbial membrane over that of the host cell is one of the most important characteristics of the antimicrobial peptides (AMPs). This selectivity strongly depends on the chemical and structural properties of the lipids that make up the cell membrane. A systematic study of the initial membrane selectivity of protegrin-1 (PG-1), a β-sheet AMP, was performed using Langmuir monolayers. Constant pressure insertion assay was used to quantify the amount of PG-1 insertion and fluorescence microscopy was employed to observe the effect of PG-1 on lipid ordering. Charge and packing properties of the monolayer were altered by using lipids with different head groups, substituting saturated with unsaturated lipid tail group(s) and incorporating spacer molecules. PG-1 inserted most readily into anionic films composed of phosphatidylglycerol (PG) and lipid A, consistent with its high selectivity for microbial membranes. It also discriminated between zwitteranionic phospholipids, inserting more readily into phosphatidylcholine (PC) monolayers than those composed of phosphatidylethanolamine, potentially explaining why PG-1 is hemolytic for PC-rich human erythrocytes and not for the PE-rich erythrocytes of ruminants. Increased packing density of the monolayer by increased surface pressure, increased tail group saturation or incorporation of dihydrocholesterol diminishes the insertion of PG-1. Fluorescence microscopy shows that lipid packing is disordered upon PG-1 insertion. However, the presence of PG-1 can still affect lipid morphology even with no observed PG-1 insertion. These results show the important role that lipid composition of the cell membrane plays in the activity of AMPs.  相似文献   

16.
Obtaining highly specific and active ribonuclease activities is an important goal with numerous medical and biochemical applications. As a step toward more active and specific ribonucleases, we describe the preparation and the enzymatic and structural properties of RNase S monomers and dimers conjugated to DNA and PNA molecules. Poly(dT)n (2'-oligodeoxyribonucleotides, n = 8, 15) and t8 peptide nucleic acid (PNA) chains have been conjugated to the S-peptide of ribonuclease S. Monomers and dimers of the conjugated enzyme have been obtained and characterized by 1H NMR spectroscopy, showing that DNA or PNA conjugation does not alter the native structure of ribonuclease S. The oligonucleotide-conjugated RNase S monomer and dimer show significant activity against single-stranded RNA and very low/negligible hydrolysis of double-stranded poly(A).poly(U). In contrast, the t8-conjugated RNase S monomer and dimer show substantial activity against both ssRNA and dsRNA. These results highlight the importance of positive charges near but not in the active site in enhancing activity against dsRNA and reveal the promise of PNA-RNase conjugates for modulating RNase activity.  相似文献   

17.
Release of free bases from calf thymus DNA upon irradiation in aerated 0.1 mol dm-3NaClO4 at pH 7 has been measured by HPLC and shown to be markedly influenced by the presence of thiols during irradiation. The ability of thiols to protect DNA was shown to depend upon the net charge (Z) at pH 7 in the order WR 1065 (Z = +2) greater than cysteamine (Z = +1) greater than 2-mercaptoethanol (Z = 0) approximately equal to dithiothreitol (Z = 0) greater than GSH (Z = -1) approximately equal to 2-mercaptoethanesulfonic acid (Z = -1) approximately equal to 2-mercaptosuccinate (Z = -2). A similar dependence of protection upon net charge was found for disulfides: cystamine (Z = +2) greater than 2-mercaptoethyl disulfide (Z = 0) greater than GSSG (Z = -2). Protection by WR 1065, but not by 2-mercaptoethanol or GSH, was found to decrease significantly with increasing ionic strength. Protection by WR 1065 and GSH was not markedly dependent upon pH between pH 6 and 8. The results are explained in terms of electrostatic interaction of the thiols with DNA, leading to high concentrations of cations near DNA, which allow them to scavenge hydroxyl radicals and repair DNA radicals effectively and to low concentrations of anionic thiols near DNA, which limit their effectiveness as protectors. Poly(dG,dC) and calf thymus DNA exhibited comparable release of G and C upon changing from 0.1 to 0.7 mol dm-3 MgSO4. Since this change causes poly(dG,dC), but not calf thymus DNA, to undergo a change from the B-form to the Z-form of DNA, both forms must have a comparable susceptibility to radiation-induced base release.  相似文献   

18.
In this study, we have examined the membrane properties and sterol interactions of phosphatidyl alcohols varying in the size of the alcohol head group coupled to the sn-3-linked phosphate. Phosphatidyl alcohols of interest were dipalmitoyl derivatives with methanol (DPPMe), ethanol (DPPEt), propanol (DPPPr), or butanol (DPPBu) head groups. The Phosphatidyl alcohols are biologically relevant, because they can be formed in membranes by the phospholipase D reaction in the presence of alcohol. The melting behavior of pure phosphatidyl alcohols and mixtures with 1,2-dipalmitoyl-sn-glycero-3-phosphocholine (DPPC) or cholesterol was assessed using high sensitivity differential scanning calorimetry (DSC). DPPMe had the highest melting temperature (∼ 49 °C), whereas the other phosphatidyl alcohols had similar melting temperatures as DPPC (∼ 40-41 °C). All phosphatidyl alcohols, except DPPMe, also showed good miscibility with DPPC. The effects of cholesterol on the melting behavior and membrane order in multilamellar bilayer vesicles were assessed using steady-state anisotropy of 1,6-diphenyl-1,3,5-hexatriene (DPH) and DSC. The ordering effect of cholesterol in the fluid phase was lower for all phosphatidyl alcohols as compared to DPPC and decreased with increasing head group size. The formation of ordered domains containing the phosphatidyl alcohols in complex bilayer membranes was determined using fluorescence quenching of DPH or the sterol analogue cholesta-5,7,(11)-trien-3-beta-ol (CTL). The phosphatidyl alcohols did not appear to form sterol-enriched ordered domains, whereas DPPMe, DPPEt appeared to form ordered domains in the temperature window examined (10-50 °C). The partitioning of CTL into bilayer membranes containing phosphatidyl alcohols was to a small extent increased for DPPMe and DPPEt, but in general, sterol interactions were weak or unfavorable for the phosphatidyl alcohols. Our results show that the biophysical and sterol interacting properties of phosphatidyl alcohols, having identical acyl chain structures, are markedly dependent on the size of the head group.  相似文献   

19.
Candidates for a subunit vaccine against bovine babesiosis include surface proteins of infective forms found in the salivary glands of tick vectors. However, low numbers of infective forms are present within ticks and hinder analysis of this stage. To solve this problem, conditions which yield high numbers of infective forms were investigated with the use of a Babesia bigemina-specific DNA probe. DNA from progeny of female Boophilus microplus infected with B. bigemina was hybridized to probe DNA to detect and quantitate infection. There was no difference in the prevalence of infection in progeny of three strains of Bo. microplus. However, within a strain, prevalence could be increased to 30% by combining selection of progeny from heavily (3+) infected female ticks and selection of eggs laid 120 hr postengorgement. Quantitation of infective forms within pooled salivary gland preparations of 10 infected nymphal and adult Bo. microplus demonstrated that Day 9 and 10 nymphal ticks contained the highest numbers of parasites and represented approximately 10(6) infective forms. This number of infective forms is suitable for isolation and further characterization.  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