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.
DNA hairpins produce ionic current signatures when captured by the alpha-hemolysin nano-scale pore under conditions of single molecule electrophoresis. Gating patterns produced by individual DNA hairpins when captured can be used to distinguish differences of a single base pair or even a single nucleotide [Vercoutere,W.A. et al. (2003) Nucleic Acids Res., 31, 1311–1318]. Here we investigate the mechanism(s) that may account for the ionic current gating signatures. The ionic current resistance profile of conductance states produced by DNA hairpin molecules with 3–12 bp stems showed a plateau in resistance between 10 and 12 bp, suggesting that hairpins with 10–12 bp stems span the pore vestibule. DNA hairpins with 9–12 bp stems produced gating signatures with the same relative conductance states. Systematic comparison of the conductance state dwell times and apparent activation energies for a series of 9–10 bp DNA hairpins suggest that the 3′ and 5′ ends interact at or near the limiting aperture within the vestibule of the alpha-hemolysin pore. The model presented may be useful in predicting and interpreting DNA detection using nanopore detectors. In addition, this well-defined molecular system may prove useful for investigating models of ligand-gated channels in biological membranes.  相似文献   

2.
Disruption of Epstein – Barr virus (EBV) latency is mediated through the activation of the viral immediate-early proteins, BZLF1 (Z) and BRLF1 (R).i.; (Chevallier-Greco, A., et al., (1986) EMBO J., 5, 3243 – 9; Countryman, and Miller, G. (1985) Proc. Natl. Acad. Sci. USA, 82, 4085 – 4089). We have previously demonstrated that these proteins cooperatively activate the EBV early promoter BMRF1 in lymphoid cells but not in epithelial cells. Although cooperative transactivation by these proteins has been demonstrated with a number of EBV promoters, the mechanism of this interaction is not well understood. We now show that the cooperative activation of the BMRF1 promoter by Z-plus-R requires an intact R binding site and at least one functional Z response element (ZRE). Despite the presence of an R binding site, the BMRF1 promoter is only moderately responsive to R alone in either HeLa or Jurkat cells. Efficient activation of the BMRF1 promoter by Z alone in HeLa cells requires two ZREs (located at − 59 and − 106), whereas two additional Z binding sites (located at − 42 and − 170) contribute very little to Z-induced activation. In the absence of ZREs, Z acted as a repressor of R-induced transactivation. These observations, along with observations made by other investigators (Giot, J.F. et al., (1991) Nucleic Acids Res., 19, 1251 – 8), suggest that Z-plus-R cooperative activation is dependent upon 1) direct binding by R and Z to responsive promoter elements and 2) contributions by cell-specific factors.  相似文献   

3.
4.
5.
The ever increasing rate at which whole genome sequences are becoming accessible to the scientific community has created an urgent need for tools enabling comparison of chromosomes of different species. We have applied biometric methods to available chromosome sequences and posted the results on our Comparative Genometrics (CG) web site. By genometrics, a term coined by Elston and Wilson [Genet. Epidemiol. (1990), 7, 17–19], we understand a biometric analysis of chromosomes. During the initial phase, our web site displays, for all completely sequenced prokaryotic genomes, three genometric analyses: the DNA walk [Lobry (1999) Microbiology Today, 26, 164–165] and two complementary representations, i.e. the cumulative GC- and TA-skew analyses, capable of identifying, at the level of whole genomes, features inherent to chromosome organization and functioning. It appears that the latter features are taxon-specific. Although primarily focused on prokaryotic chromosomes, the CG web site contains genometric information on paradigm plasmids, phages, viruses and eukaryotic organelles. Relevant data and methods can be readily used by the scientific community for further analyses as well as for tutorial purposes. Our data posted at the CG web site are freely available on the World Wide Web at http://www.unil.ch/comparativegenometrics.  相似文献   

6.
Most studies on the structure of DNA in telomeres have been dedicated to the double-stranded region or the guanosine-rich strand and consequently little is known about the factors that may bind to the telomere cytosine-rich (C-rich) strand. This led us to investigate whether proteins exist that can recognise C-rich sequences. We have isolated several nuclear factors from human cell extracts that specifically bind the C-rich strand of vertebrate telomeres [namely a d(CCCTAA)n repeat] with high affinity and bind double-stranded telomeric DNA with a 100× reduced affinity. A biochemical assay allowed us to characterise four proteins of apparent molecular weights 66–64, 45 and 35 kDa, respectively. To identify these polypeptides we screened a λgt11-based cDNA expression library, obtained from human HeLa cells using a radiolabelled telomeric oligonucleotide as a probe. Two clones were purified and sequenced: the first corresponded to the hnRNP K protein and the second to the ASF/SF2 splicing factor. Confirmation of the screening results was obtained with recombinant proteins, both of which bind to the human telomeric C-rich strand in vitro.  相似文献   

7.
Kim SJ  Kim BH 《Nucleic acids research》2003,31(11):2725-2734
We have synthesized three types of calix[4]arene– nucleoside hybrid efficiently by amide bond formation between the amine functional groups of 1,3-diaminocalix[4]arene and the carboxyl groups of thymidine nucleoside derivatives. X-ray crystallography of a homocoupled calix[4]arene–nucleoside hybrid revealed an interesting hydrogen bonding pattern between thymine bases and the amide linkages. We designed the calix[4]arene–oligonucleotide hybrids (5′-AAAAGATATCAAXTTGATATCTTTT-3′, 5′-T12-X-T12-3′, and 5′-A12-X-T12-3′) to be V-shaped oligodeoxyribonucleotides and synthesized them by using a calix[4]arene–nucleoside hybrid (X) as a key building block. Thermal denaturation experiments, monitored by UV spectroscopy at 260 and 284 nm, and circular dichroism spectra of the calix[4]arene–oligonucleotide hybrids suggest that the modified oligonucleotides indeed adopt V-shaped conformations, making them suitable for use as building blocks in the construction of programmed oligonucleotide nanostructures.  相似文献   

8.
Wu J  Zhu BB  Yu J  Zhu H  Qiu L  Kindy MS  Gu L  Seidel A  Li GM 《Nucleic acids research》2003,31(22):6428-6434
Benzo[c]phenanthrene dihydrodiol epoxide (B[c] PhDE) is well known as an important environmental chemical carcinogen that preferentially modifies DNA in adenine residues. However, the molecular mechanism by which B[c]PhDE induces tumorigenesis is not fully understood. In this report, we demonstrate that DNA mismatch repair (MMR), a genome maintenance system, plays an important role in B[c]PhDE-induced carcinogensis by promoting apoptosis in cells treated with B[c]PhDE. We show that purified human MMR recognition proteins, MutSα and MutSβ, specifically recognized B[c]PhDE-DNA adducts. Cell lines proficient in MMR exhibited several-fold more sensitivity to killing than cell lines defective in either MutSα or MutLα by B[c]PhDE; the nature of this sensitivity was shown to be due to increased apoptosis. Additionally, wild-type mice exposed to B[c]PhDE had intestinal crypt cells that underwent apoptosis significantly more often than intestinal crypt cells found in B[c]PhDE-treated Msh2–/– or Mlh1–/– mice. These findings, combined with previous studies, suggest that the MMR system may serve as a general sensor for chemical-caused DNA damage to prevent damaged cells from mutagenesis and carcinogenesis by promoting apoptosis.  相似文献   

9.
Enzymes that form transient DNA–protein covalent complexes are targets for several potent classes of drugs used to treat infectious disease and cancer, making it important to establish robust and rapid procedures for analysis of these complexes. We report a method for isolation of DNA–protein adducts and their identification and quantification, using techniques compatible with high-throughput screening. This method is based on the RADAR assay for DNA adducts that we previously developed (Kiianitsa and Maizels (2013) A rapid and sensitive assay for DNA–protein covalent complexes in living cells. Nucleic Acids Res., 41:e104), but incorporates three key new steps of broad applicability. (i) Silica-assisted ethanol/isopropanol precipitation ensures reproducible and efficient recovery of DNA and DNA–protein adducts at low centrifugal forces, enabling cell culture and DNA precipitation to be carried out in a single microtiter plate. (ii) Rigorous purification of DNA–protein adducts by a procedure that eliminates free proteins and free nucleic acids, generating samples suitable for detection of novel protein adducts (e.g. by mass spectroscopy). (iii) Identification and quantification of DNA–protein adducts by direct ELISA assay. The ELISA-based RADAR assay can detect Top1–DNA and Top2a–DNA adducts in human cells, and gyrase–DNA adducts in Escherichia coli. This approach will be useful for discovery and characterization of new drugs to treat infectious disease and cancer, and for development of companion diagnostics assays for individualized medicine.  相似文献   

10.
The ability of Enterococcus faecalis to metabolically adapt to an oligotrophic environment has been analyzed. E. faecalis is able to survive for prolonged periods under conditions of complete starvation established by incubation in tap water. During incubation in this microcosm, cells developed a rippled cell surface with irregular shapes. Exponentially growing cells survived to the same extent as cells starved for glucose prior to exposure to the multiple nutrient deficient stress. Chloramphenicol treatment during incubation in tap water led to a rapid decline in plate counts for exponentially growing cells but showed progressively reduced influence on stationary-phase cells harvested after different times of glucose starvation. During incubation in the oligotrophic environment, cells from the exponential-growth phase and early-stationary phase became progressively more resistant to other environmental stresses (heat [62°C], acid [pH 3.3], UV254 nm light [180 J/m2], and sodium hypochlorite [0.05%]) until they reached a maximum of survival characteristic for each treatment. In contrast, cells starved of glucose for 24 h did not become more resistant to the different treatments during incubation in tap water. Our combined data suggest that energy starvation induces a response similar to that triggered by oligotrophy. Analysis of protein synthesis by two-dimensional gel electrophoresis revealed the enhanced synthesis of 51 proteins which were induced in the oligotrophic environment. A comparison of these oligotrophy-inducible proteins with the 42 glucose starvation-induced polypeptides (J. C. Giard, A. Hartke, S. Flahaut, P. Boutibonnes, and Y. Auffray, Res. Microbiol. 148:27–35, 1997) showed that 16 are common between the two different starvation conditions. These proteins and the corresponding genes seem to play a key role in the observed phenomena of long-term survival and development of general stress resistance of starved cultures of E. faecalis.  相似文献   

11.
Oligonucleotides incorporating 8-aza-7-deazapurin-2,6-diamine (pyrazolo[3,4-d]pyrimidin-4,6-diamine) nucleoside 2a or its 7-bromo derivative 2b show enhanced duplex stability compared to those containing dA. While incorporation of 2a opposite dT increases the Tm value only slightly, the 7-bromo compound 2b forms a very stable base pair which is as strong as the dG-dC pair. Compound 2b shows a similar base discrimination in duplex DNA as dA. The base-modified nucleosides 2a,b have a significantly more stable N-glycosylic bond than the rather labile purin-2,6-diamine 2′-deoxyribonucleoside 1. Base protection with acyl groups, with which we had difficulties in the case of purine nucleoside 1, was effective with pyrazolo[3,4-d]-pyrimidine nucleosides 2a,b. Oligonucleotides containing 2a,b were obtained by solid phase synthesis employing phosphoramidite chemistry. Compound 2b harmonizes the stability of DNA duplexes. Their stability is no longer dependent on the base pair composition while they still maintain their sequence specificity. Thus, they have the potential to reduce the number of mispairs when hybridized in solution or immobilized on arrays.  相似文献   

12.
A key issue in the nucleotide excision repair (NER) of bulky carcinogen–DNA adducts is the ability of the NER machinery to recognize and repair certain adducts while failing to repair others. Unrepaired adducts can survive to cause mutations that initiate the carcinogenic process. Benzo[c]phenanthrene (B[c]Ph), a representative fjord region polycyclic aromatic hydrocarbon, can be metabolically activated to the enantiomeric benzo[c]phenanthrene diol epoxides (B[c]PhDEs), (+)-(1S,2R,3R,4S)-3,4- dihydroxy-1,2-epoxy-1,2,3,4-tetrahydrobenzo[c]phe nanthrene and the corresponding (–)-(1R,2S,3S,4R) isomer. These react predominantly with adenine residues in DNA to produce the stereoisomeric 1R (+)- and 1S (–)-trans-anti-B[c]Ph-N6-dA adducts. Duplexes containing the 1R (+) or 1S (–) B[c]Ph-dA adduct in codon 61 of the human N-ras mutational hotspot sequence CA*A, with B[c]Ph modification at A*, are not repaired by the human NER system. However, the analogous stereoisomeric DNA adducts of the bay region benzo[a]pyrene diol epoxide (B[a]PDE), 10S (+)- and 10R (–)-trans-anti-B[a]P-N6-dA, are repaired in the same base sequence. In order to elucidate structural and thermodynamic origins of this phenomenon, we have carried out a 2 ns molecular dynamics simulation for the 1R (+)- and 1S (–)-trans-anti-B[c]Ph-N6-dA adducts in an 11mer duplex containing the human N-ras codon 61 sequence, and compared these results with our previous study of the B[a]P-dA adducts in the same sequence. The molecular mechanics Poisson– Boltzmann surface area (MM-PBSA) method was applied to calculate the free energies of the pair of stereoisomeric B[c]Ph-dA adducts, and a detailed structural analysis was carried out. The different repair susceptibilities of the B[a]P-dA adducts and the B[c]Ph-dA adducts can be attributed to different degrees of distortion, stemming from combined effects of differences in the quality of Watson–Crick hydrogen bonding, unwinding, stretching and helix backbone perturbations. These differences are due to the different intrinsic topologies of the rigid, planar bay region adducts versus the twisted, sterically hindered fjord region adducts.  相似文献   

13.
We report on the physicochemical effects resulting from incorporating a 5-(3-aminopropyl) side chain onto a 2′-deoxyuridine (dU) residue in a short DNA hairpin. A combination of spectroscopy, calorimetry, density and ultrasound techniques were used to investigate both the helix–coil transition of a set of hairpins with the following sequence: d(GCGACTTTTTGNCGC) [N = dU, deoxythymidine (dT) or 5-(3-aminopropyl)-2′-deoxyuridine (dU*)], and the interaction of each hairpin with Mg2+. All three molecules undergo two-state transitions with melting temperatures (TM) independent of strand concentration that indicates their intramolecular hairpin formation. The unfolding of each hairpin takes place with similar TM values of 64–66°C and similar thermodynamic profiles. The unfavorable unfolding free energies of 6.4–6.9 kcal/mol result from the typical compensation of unfavorable enthalpies, 36–39 kcal/mol, and favorable entropies of ~110 cal/mol. Furthermore, the stability of each hairpin increases as the salt concentration increases, the TM-dependence on salt yielded slopes of 2.3–2.9°C, which correspond to counterion releases of 0.53 (dU and dT) and 0.44 (dU*) moles of Na+ per mole of hairpin. Absolute volumetric and compressibility measurements reveal that all three hairpins have similar hydration levels. The electrostatic interaction of Mg2+ with each hairpin yielded binding affinities in the order: dU > dT > dU*, and a similar release of 2–4 electrostricted water molecules. The main result is that the incorporation of the cationic 3-aminopropyl side chain in the major groove of the hairpin stem neutralizes some local negative charges yielding a hairpin molecule with lower charge density.  相似文献   

14.
The core protein of hepatitis C virus (HCV) has been shown previously to act as a potent nucleic acid chaperone in vitro, promoting the dimerization of the 3′-untranslated region (3′-UTR) of the HCV genomic RNA, a process probably mediated by a small, highly conserved palindromic RNA motif, named DLS (dimer linkage sequence) [G. Cristofari, R. Ivanyi-Nagy, C. Gabus, S. Boulant, J. P. Lavergne, F. Penin and J. L. Darlix (2004) Nucleic Acids Res., 32, 2623–2631]. To investigate in depth HCV RNA dimerization, we generated a series of point mutations in the DLS region. We find that both the plus-strand 3′-UTR and the complementary minus-strand RNA can dimerize in the presence of core protein, while mutations in the DLS (among them a single point mutation that abolished RNA replication in a HCV subgenomic replicon system) completely abrogate dimerization. Structural probing of plus- and minus-strand RNAs, in their monomeric and dimeric forms, indicate that the DLS is the major if not the sole determinant of UTR RNA dimerization. Furthermore, the N-terminal basic amino acid clusters of core protein were found to be sufficient to induce dimerization, suggesting that they retain full RNA chaperone activity. These findings may have important consequences for understanding the HCV replicative cycle and the genetic variability of the virus.  相似文献   

15.
Primer extension studies have shown that the Y-family DNA polymerase IV (Dpo4) from Sulfolobus solfataricus P2 can preferentially insert C opposite N-(deoxyguanosin-8-yl)-2-acetylaminofluorene (AAF-dG) [F. Boudsocq, S. Iwai, F. Hanaoka and R. Woodgate (2001) Nucleic Acids Res., 29, 4607–4616]. Our goal is to elucidate on a structural level how AAF-dG can be harbored in the Dpo4 active site opposite an incoming dCTP, using molecular modeling and molecular dynamics simulations, since AAF-dG prefers the syn glycosidic torsion. Both anti and syn conformations of the templating AAF-dG in a Dpo4 ternary complex were investigated. All four dNTPs were studied. We found that an anti glycosidic torsion with C1′-exo deoxyribose conformation allows AAF-dG to be Watson–Crick hydrogen-bonded with dCTP with modest polymerase perturbation, but other nucleotides are more distorting. The AAF is situated in the Dpo4 major groove open pocket with fluorenyl rings 3′- and acetyl 5′-directed along the modified strand, irrespective of dNTP. With AAF-dG syn, the fluorenyl rings are in the small minor groove pocket and the active site region is highly distorted. The anti-AAF-dG conformation with C1′-exo sugar pucker can explain the preferential incorporation of dC by Dpo4. Possible relevance of our new major groove structure for AAF-dG to other polymerases, lesion repair and solution conformations are discussed.  相似文献   

16.
Many nucleic acid binding proteins use short peptide sequences to provide specificity in recognizing their targets, which may be either a specific sequence or a conformation. Peptides containing alternating lysine have been shown to bind to poly(dG–d5meC) in the Z conformation, and stabilize the higher energy form [H. Takeuchi, N. Hanamura, H. Hayasaka and I. Harada (1991) FEBS Lett., 279, 253–255 and H. Takeuchi, N. Hanamura and I. Harada (1994) J. Mol. Biol., 236, 610–617.]. Here we report the construction of a Z-DNA specific binding protein, with the peptide KGKGKGK as a functional domain and a leucine zipper as a dimerization domain. The resultant protein, KGZIP, induces the Z conformation in poly(dG–d5meC) and binds to Z-DNA stabilized by bromination with high affinity and specificity. The binding of KGZIP is sufficient to convert poly(dG–d5meC) from the B to the Z form, as shown by circular dichroism. The sequence KGKGKGK is found in many proteins, although no functional role has been established. KGZIP also has potential for engineering other Z-DNA specific proteins for future studies of Z-DNA in vitro and in vivo.  相似文献   

17.
The biophysical properties of oligodeoxyribonucleotides (ODNs) selectively modified with conformationally ‘locked’ bicyclo[3.1.0]hexane pseudosugars (Maier,M.A., Choi,Y., Gaus,H., Barchi,J.J. Jr, Marquez,V.E., Manoharan,M. (2004) Synthesis and characterization of oligonucleotides containing conformationally constrained bicyclo[3.1.0]hexane pseudosugar analogs Nucleic Acids Res., 32, 3642–3650) have been studied by various techniques. Six separate synthetic ODNs based on the Dickerson Drew dodecamer sequence (CGCGAAT*T*CGCG) were examined where each one (or both) of the thymidines (T*) were substituted with a bicyclic pseudosugar locked in either a North (2′-exo) or South (3′-exo) ring pucker. Circular dichroism spectroscopy, differential scanning calorimetry and 1H NMR spectroscopy were used to examine the duplex stability and conformational properties of the ODNs. Replacement of one or both thymidines with North-locked sugars (RNA-like) into the dodecamer did not greatly affect duplex formation or melt temperatures but distinct differences in thermodynamic parameters were observed. In contrast, incorporation of South-locked sugar derivatives that were predicted to stabilize this standard B-DNA, had the unexpected effect of causing a conformational equilibrium between different duplex forms at specific strand and salt concentrations. Our data and those of others suggest that although DNA can tolerate modifications with RNA-like (North) nucleotides, a more complicated spectrum of changes emerges with modifications restricted to South (DNA-like) puckers.  相似文献   

18.
RNA-binding properties of the mitochondrial Y-box protein RBP16   总被引:6,自引:3,他引:6  
We have previously identified a mitochondrial Y-box protein in Trypanosoma brucei that we designated RBP16. The predicted RBP16 amino acid sequence revealed the presence of a cold-shock domain at its N-terminus and a glycine- and arginine-rich C-terminus reminiscent of an RGG RNA-binding motif. Since RBP16 is capable of interacting with different guide RNAs (gRNAs) in vitro and in vivo primarily via the oligo(U) tail, as well as with ribosomal RNAs, possible functions of RBP16 may be in kinetoplastid RNA editing and/or translation. Herein, we report experiments that further define the RNA-binding properties of RBP16. RBP16 forms a single stable complex with the gRNA gA6[14] at low protein concentration, while at higher protein concentration two stable complexes that possibly represent two different conformations are observed. Both complexes are stable at relatively high salt and moderate heparin concentrations indicating that the binding of RBP16 to gA6[14] does not rely primarily on ionic interactions. Phenylglyoxal treatment of the protein indicates that arginine residues are important in RNA binding. The minimal length of RNA sequence necessary for the binding of RBP16 was assessed by gel retardation and UV cross-linking competition assays using oligo(U) ribonucleotides of varying lengths (4–40 nt). Although RBP16 can bind to oligonucleotides as small as U4, its affinity increases with the length of the oligo(U) ribonucleotide, with a dramatic increase in binding efficiency observed when the length is increased to 10 nt. Gel retardation assays employing T.brucei mRNAs demonstrated that, although it acts as a major binding determinant, a 3′ U tail is not an absolute requirement for efficient RBP16RNA binding. Experiments with oligonucleotides containing U stretches embedded at different positions in oligo(dC) indicated that high-affinity binding requires both a uridine stretch, as well as 5′ and 3′ non-specific sequences. These results suggest a model for the molecular interactions involved in RBP16RNA binding.  相似文献   

19.
The solution structure and hydration of the chimeric duplex [d(CGC)r(aaa)d(TTTGCG)]2, in which the central hybrid segment is flanked by DNA duplexes at both ends, was determined using two-dimensional NMR, simulated annealing and restrained molecular dynamics. The solution structure of this chimeric duplex differs from the previously determined X-ray structure of the analogous B-DNA duplex [d(CGCAAATTTGCG)]2 as well as NMR structure of the analogous A-RNA duplex [r(cgcaaauuugcg)]2. Long-lived water molecules with correlation time τc longer than 0.3 ns were found close to the RNA adenine H2 and H1′ protons in the hybrid segment. A possible long-lived water molecule was also detected close to the methyl group of 7T in the RNA–DNA junction but not with the other two thymines (8T and 9T). This result correlates with the structural studies that only DNA residue 7T in the RNA–DNA junction adopts an O4′-endo sugar conformation, while the other DNA residues including 3C in the DNA–RNA junction, adopt C1′-exo or C2′-endo conformations. The exchange rates for RNA C2′-OH were found to be ~520 s–1. This slow exchange rate may be due to the narrow minor groove width of [d(CGC)r(aaa)d(TTTGCG)]2, which may trap the water molecules and restrict the dynamic motion of hydroxyl protons. The minor groove width of [d(CGC)r(aaa)d(TTTGCG)]2 is wider than its B-DNA analog but narrower than that of the A-RNA analog. It was further confirmed by its titration with the minor groove binding drug distamycin. A possible 2:1 binding mode was found by the titration experiments, suggesting that this chimeric duplex contains a wider minor groove than its B-DNA analog but still narrow enough to hold two distamycin molecules. These distinct structural features and hydration patterns of this chimeric duplex provide a molecular basis for further understanding the structure and recognition of DNA·RNA hybrid and chimeric duplexes.  相似文献   

20.
Thermodynamically based DNA strand design   总被引:3,自引:0,他引:3       下载免费PDF全文
We describe a new algorithm for design of strand sets, for use in DNA computations or universal microarrays. Our algorithm can design sets that satisfy any of several thermodynamic and combinatorial constraints, which aim to maximize desired hybridizations between strands and their complements, while minimizing undesired cross-hybridizations. To heuristically search for good strand sets, our algorithm uses a conflict-driven stochastic local search approach, which is known to be effective in solving comparable search problems. The PairFold program of Andronescu et al. [M. Andronescu, Z. C. Zhang and A. Condon (2005) J. Mol. Biol., 345, 987–1001; M. Andronescu, R. Aguirre-Hernandez, A. Condon, and H. Hoos (2003) Nucleic Acids Res., 31, 3416–3422.] is used to calculate the minimum free energy of hybridization between two mismatched strands. We describe new thermodynamic measures of the quality of strand sets. With respect to these measures of quality, our algorithm consistently finds, within reasonable time, sets that are significantly better than previously published sets in the literature.  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