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.
Ensemble and single-molecule fluorescence measurements of 2′-(4-hydroxyphenyl)-5-[5-(4-methylpiperazine-1-yl) benzimidazo-2-yl]-benzimidazole (H-258)– calf thymus (CT) DNA complexes at various [H-258]/[DNA bp] ratios were performed to elucidate the binding of H-258 with DNA. Upon binding to double-stranded CT DNA (CT ds DNA) at a [H-258]/[DNA bp] ratio of 0.05 the relative fluorescence quantum yield, Φf, of H-258 increases from 0.02 to 0.58. The fluorescence decay can be fitted almost by a mono-exponential model with a lifetime of ~3.6 ns. This indicates that H-258 binds almost quantitatively in the minor groove of DNA at low [H-258]/[DNA bp] ratios. With increasing [H-258]/[DNA bp] ratios, e.g. 0.15 and 0.20, the fluorescence quantum yield of H-258 decreases to 0.28 and 0.19, respectively. Fitting of the fluorescence decays measured for higher [H-258]/[DNA bp] ratios reveals the presence of additional shorter fluorescence lifetime components in the range of 0.5–2.0 ns. Our results suggest that H-258 partially intercalates in G:C sequences at higher [H-258]/[DNA bp] ratios reflected by a lifetime component of 1.5–2 ns. In addition, stacking or adsorption of H-258 molecules on DNA occurs at higher [H-258]/[DNA bp] ratios. These molecules exhibit a short fluorescence lifetime of ~500 ps and are more exposed to the aqueous environment. Fluorescence transients of the intensity and lifetime of single H-258 CT ds DNA demonstrate that weakly (unspecific) bound H-258 molecules exhibit a shorter fluorescence lifetime and a strongly reduced photostability.  相似文献   

2.
A new asymmetric cyanine dye has been synthesised and its interaction with different DNA has been investigated. In this dye, BEBO, the structure of the known intercalating cyanine dye BO has been extended with a benzothiazole substituent. The resulting crescent-shape of the molecule is similar to that of the well-known minor groove binder Hoechst 33258. Indeed, comparative studies of BO illustrate a considerable change in binding mode induced by this structural modification. Linear and circular dichroism studies indicate that BEBO binds in the minor groove to [poly (dA-dT)](2), but that the binding to calf thymus DNA is heterogeneous, although still with a significant contribution of minor groove binding. Similar to other DNA binding asymmetric cyanine dyes, BEBO has a large increase in fluorescence intensity upon binding and a relatively large quantum yield when bound. The minor groove binding of BEBO to [poly (dA-dT)](2) affords roughly a 180-fold increase in intensity, which is larger than to that of the commonly used minor groove binding probes DAPI and Hoechst 33258.  相似文献   

3.
The detection of double-stranded (ds) DNA by SYBR Green I (SG) is important in many molecular biology methods including gel electrophoresis, dsDNA quantification in solution and real-time PCR. Biophysical studies at defined dye/base pair ratios (dbprs) were used to determine the structure–property relationships that affect methods applying SG. These studies revealed the occurrence of intercalation, followed by surface binding at dbprs above ~0.15. Only the latter led to a significant increase in fluorescence. Studies with poly(dA) · poly(dT) and poly(dG) · poly(dC) homopolymers showed sequence-specific binding of SG. Also, salts had a marked impact on SG fluorescence. We also noted binding of SG to single-stranded (ss) DNA, although SG/ssDNA fluorescence was at least ~11-fold lower than with dsDNA. To perform these studies, we determined the structure of SG by mass spectrometry and NMR analysis to be [2-[N-(3-dimethylaminopropyl)-N-propylamino]-4-[2,3-dihydro-3-methyl-(benzo-1,3-thiazol-2-yl)-methylidene]-1-phenyl-quinolinium]. For comparison, the structure of PicoGreen (PG) was also determined and is [2-[N-bis-(3-dimethylaminopropyl)-amino]-4-[2,3-dihydro-3-methyl-(benzo-1,3-thiazol-2-yl)-methylidene]-1-phenyl-quinolinium]+. These structure–property relationships help in the design of methods that use SG, in particular dsDNA quantification in solution and real-time PCR.  相似文献   

4.
High precision ultrasonic and densimetric techniques have been used to study the interaction of Ni2+ ions with right-handed poly[d(G-C)]·poly[d(G-C)], poly-[d(A-C)]·poly[d(G-T)] and poly[d(A-T)]·poly[d(A-T)] in 5 mM CsCl, 0.2 mM HEPES, pH 7.5 at 20°C. From these measurements the changes in the apparent molar volume and the apparent molar adiabatic compressibility due to the interaction have been obtained. The volume effects of the binding, calculated per mole of Ni2+ ions, range from 11.7 to 23.9 cm3 mol–1 and the compressibility effects range from 19.3 × 10–4 to 43.1 × 10–4 cm3 mol–1 bar–1. These data are interpreted in terms of dehydration of the polynucleotides and Ni2+ ions, i.e. the release of water molecules from the hydration shells of the molecules. An increase in G+C content gives an increase in volume and compressibility effects, indicating a rise in the extent of dehydration. The dehydration effects of Ni2+ binding to poly[d(G-C)]·poly[d(G-C)] are approximately twice those of poly[d(A-T)]·poly[d(A-T)]. The volume and compressibility effects of Ni2+–EDTA complex formation have also been measured and used as a model system for quantitative estimation. These values revealed that Ni2+ ions can coordinate two atomic groups of poly[d(G-C)]·poly[d(G-C)], while in the case of the Ni2+–poly[d(A-T)]·poly[d(A-T)] complex volume and compressibility effects correspond to one direct or two indirect (through water) contacts.  相似文献   

5.
Two new crescent-shaped unsymmetrical cyanine dyes have been synthesised and their interactions with DNA have been investigated by different spectroscopic methods. These dyes are analogues to a minor groove binding unsymmetrical cyanine dye, BEBO, recently reported by us. In this dye, the structure of the known intercalating cyanine dye BO was extended with a benzothiazole substituent. To investigate how the identity of the extending heterocycle affects the binding to DNA, the new dyes BETO and BOXTO have a benzothiazole group and a benzoxazole moiety, respectively. Whereas BEBO showed a heterogeneous binding to calf thymus DNA, linear and circular dichroism studies of BOXTO indicate a high preference for minor groove binding. BETO also binds in the minor groove to mixed sequence DNA but has a contribution of non-ordered and non-emissive species present. A non-intercalative binding mode of the new dyes, as well as for BEBO, is further supported by electrophoresis unwinding assays. These dyes, having comparable spectral properties as the intercalating cyanine dyes, but bind in the minor groove instead, might be useful complements for staining of DNA. In particular, the benzoxazole substituted dye BOXTO has attractive fluorescence properties with a quantum yield of 0.52 when bound to mixed sequence DNA and a 300-fold increase in fluorescence intensity upon binding.  相似文献   

6.
SYBR Gold is a commonly used and particularly bright fluorescent DNA stain, however, its chemical structure is unknown and its binding mode to DNA remains controversial. Here, we solve the structure of SYBR Gold by NMR and mass spectrometry to be [2-(4-{[diethyl(methyl)ammonio]methyl}phenyl)-6-methoxy-1-methyl-4-{[(2Z)-3-methyl-1,3-benzoxazol-2-ylidene]methyl}quinolin-1-ium] and determine its extinction coefficient. We quantitate SYBR Gold binding to DNA using two complementary approaches. First, we use single-molecule magnetic tweezers (MT) to determine the effects of SYBR Gold binding on DNA length and twist. The MT assay reveals systematic lengthening and unwinding of DNA by 19.1° ± 0.7° per molecule upon binding, consistent with intercalation, similar to the related dye SYBR Green I. We complement the MT data with spectroscopic characterization of SYBR Gold. The data are well described by a global binding model for dye concentrations ≤2.5 μM, with parameters that quantitatively agree with the MT results. The fluorescence increases linearly with the number of intercalated SYBR Gold molecules up to dye concentrations of ∼2.5 μM, where quenching and inner filter effects become relevant. In summary, we provide a mechanistic understanding of DNA-SYBR Gold interactions and present practical guidelines for optimal DNA detection and quantitative DNA sensing applications using SYBR Gold.  相似文献   

7.
Magnesium ions play important roles in the structure and function of nucleic acids. Whereas the tertiary folding of RNA often requires magnesium ions binding to tight places where phosphates are clustered, the molecular basis of the interactions of magnesium ions with RNA helical regions is less well understood. We have refined the crystal structures of four decamer oligonucleotides, d(ACCGGCCGGT), r(GCG)d(TATACGC), r(GC)d(GTATACGC) and r(G)d(GCGTATACGC) with bound hexahydrated magnesium ions at high resolution. The structures reveal that A-form nucleic acid has characteristic [Mg(H2O)6]2+ binding modes. One mode has the ion binding in the deep major groove of a GpN step at the O6/N7 sites of guanine bases via hydrogen bonds. Our crystallographic observations are consistent with the recent NMR observations that in solution [Co(NH3)6]3+, a model ion of [Mg(H2O)6]2+, binds in an identical manner. The other mode involves the binding of the ion to phosphates, bridging across the outer mouth of the narrow major groove. These [Mg(H2O)6]2+ ions are found at the most negative electrostatic potential regions of A-form duplexes. We propose that these two binding modes are important in the global charge neutralization, and therefore stability, of A-form duplexes.  相似文献   

8.
Genetic evidence showed two non-Mendelian genetic elements of Saccharomyces cerevisiae, called [URE3] and [PSI], to be prions of Ure2p and Sup35p, respectively. [URE3] makes cells derepressed for nitrogen catabolism, while [PSI] elevates the efficiency of weak suppressor tRNAs. The same approach led to identification of the non-Mendelian element [Het-s] of the filamentous fungus Podospora anserina, as a prion of the het-s protein. The prion form of the het-s protein is required for heterokaryon incompatibility, a normal fungal function, suggesting that other normal cellular functions may be controlled by prions. [URE3] and [PSI] involve a self-propagating aggregation of Ure2p and Sup35p, respectively. In vitro, Ure2p and Sup35p form amyloid, a filamentous protein structure, high in β-sheet with a characteristic green birefringent staining by the dye Congo Red. Amyloid deposits are a cardinal feature of Alzheimer’s disease, non-insulin-dependent diabetes mellitus, the transmissible spongiform encephalopathies, and many other diseases. The prion domain of Ure2p consists of Asn-rich residues 1 to 80, but two nonoverlapping fragments of the molecule can, when overproduced, induce the de nova appearance of [URE3]. The prion domain of Sup35 consists of residues 1 to 114, also rich in Asn and Gln residues. While runs of Asn and Gln are important for [URE3] and [PSI], no such structures are found in PrP or the Het-s protein. Either elevated or depressed levels of the chaperone Hsp104 interfere with propagation of [PSI]. Both [URE3] and [PSI] are cured by growth of cells in millimolar guanidine HCl. [URE3] is also cured by overexpression of fragments of Ure2p or fusion proteins including parts of Ure2p.  相似文献   

9.
Metal ions, and magnesium in particular, are known to be involved in RNA folding by stabilizing secondary and tertiary structures, and, as cofactors, in RNA enzymatic activity. We have conducted a systematic crystallographic analysis of cation binding to the duplex form of the HIV-1 RNA dimerization initiation site for the subtype-A and -B natural sequences. Eleven ions (K+, Pb2+, Mn2+, Ba2+, Ca2+, Cd2+, Sr2+, Zn2+, Co2+, Au3+ and Pt4+) and two hexammines [Co (NH3)6]3+ and [Ru (NH3)6]3+ were found to bind to the DIS duplex structure. Although the two sequences are very similar, strong differences were found in their cation binding properties. Divalent cations bind almost exclusively, as Mg2+, at ‘Hoogsteen’ sites of guanine residues, with a cation-dependent affinity for each site. Notably, a given cation can have very different affinities for a priori equivalent sites within the same molecule. Surprisingly, none of the two hexammines used were able to efficiently replace hexahydrated magnesium. Instead, [Co (NH3)4]3+ was seen bound by inner-sphere coordination to the RNA. This raises some questions about the practical use of [Co (NH3)6]3+ as a [Mg (H2O)6]2+ mimetic. Also very unexpected was the binding of the small Au3+ cation exactly between the Watson–Crick sites of a G-C base pair after an obligatory deprotonation of N1 of the guanine base. This extensive study of metal ion binding using X-ray crystallography significantly enriches our knowledge on the binding of middleweight or heavy metal ions to RNA, particularly compared with magnesium.  相似文献   

10.
Al toxicity is a major problem that limits crop productivity on acid soils. It has been suggested that Al toxicity is linked to changes in cellular Ca homeostasis and the blockage of plasma membrane Ca2+-permeable channels. BY-2 suspension-cultured cells of tobacco (Nicotiana tabacum L.) exhibit rapid cell expansion that is sensitive to Al. Therefore, the effect of Al on changes in cytoplasmic free Ca concentration ([Ca2+]cyt) was followed in BY-2 cells to assess whether Al perturbed cellular Ca homeostasis. Al exposure resulted in a prolonged reduction in [Ca2+]cyt and inhibition of growth that was similar to the effect of the Ca2+ channel blocker La3+ and the Ca2+ chelator ethyleneglycol-bis(β-aminoethyl ether)-N,N′-tetraacetic acid. The Ca2+ channel blockers verapamil and nifedipine did not induce a decrease in [Ca2+]cyt in these cells and also failed to inhibit growth. Al and La3+, but not verapamil or nifedipine, reduced the rate of Mn2+ quenching of Indo-1 fluorescence, which is consistent with the blockage of Ca2+- and Mn2+-permeable channels. These results suggest that Al may act to block Ca2+ channels at the plasma membrane of plant cells and this action may play a crucial role in the phytotoxic activity of the Al ion.  相似文献   

11.
[URE3] is a prion of the yeast Ure2 protein. Hsp40 is a cochaperone that regulates Hsp70 chaperone activity. When overexpressed, the Hsp40 Ydj1p cures yeast of [URE3], but the Hsp40 Sis1p does not. On the basis of biochemical data Ydj1p has been proposed to cure [URE3] by binding soluble Ure2p and preventing it from joining prion aggregates. Here, we mutagenized Ydj1p and find that disrupting substrate binding, dimerization, membrane association, or ability to transfer substrate to Hsp70 had little or no effect on curing. J-domain point mutations that disrupt functional interactions of Ydj1p with Hsp70 abolished curing, and the J domain alone cured [URE3]. Consistent with heterologous J domains possessing similar Hsp70 regulatory activity, the Sis1p J domain also cured [URE3]. We further show that Ydj1p is not essential for [URE3] propagation and that depletion of Ure2p is lethal in cells lacking Ydj1p. Our data imply that curing of [URE3] by overproduced Ydj1p does not involve direct interaction of Ydj1p with Ure2p but rather works through regulation of Hsp70 through a specific J-protein/Hsp70 interaction.  相似文献   

12.
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   

13.
The interaction of hexamminecobalt(III), Co(NH3)63+, with 160 and 3000–8000 bp length calf thymus DNA has been investigated by circular dichroism, acoustic and densimetric techniques. The acoustic titration curves of 160 bp DNA revealed three stages of interaction: (i) Co(NH3)63+ binding up to the molar ratio [Co(NH3)63+]/[P] = 0.25, prior to DNA condensation; (ii) a condensation process between [Co(NH3)63+]/[P] = 0.25 and 0.30; and (iii) precipitation after [Co(NH3)63+]/[P] = 0.3. In the case of 3000–8000 bp DNA only two processes were observed: (i) binding up to [Co(NH3)63+]/[P] = 0.3; and (ii) precipitation after this point. In agreement with earlier observations, long DNA aggregates without changes in its B-form circular dichroism spectrum, while short DNA demonstrates a positive B→Ψ transition after [Co(NH3)63+]/[P] = 0.25. From ultrasonic and densimetric measurements the effects of Co(NH3)63+ binding on volume and compressibility have been obtained. The binding of Co(NH3)63+ to both short and long DNA is characterized by similar changes in volume and compressibility calculated per mole Co(NH3)63+: ΔV = 9 cm3 mol–1 and Δκ = 33 × 10–4 cm3 mol–1 bar–1. The positive sign of the parameters indicates dehydration, i.e. water release from Co(NH3)63+ and the atomic groups of DNA. This extent of water displacement would be consistent with the formation of two direct, hydrogen bonded contacts between the cation and the phosphates of DNA.  相似文献   

14.
Molecular chaperones of the 70 kDa family mediate protein–protein interactions by selectively binding to partially unfolded segments of other proteins in an ATP-dependent activity cycle. Previous investigations of chaperone substrate selectivity have shown that chaperones have a propensity to bind to partially unfolded segments of polypeptides that contain bulky hydrophobic residues. However, recent investigations have shown that 70 kDa chaperones such as DnaK, which is expressed by Escherichia coli, also bind short basic peptides and even polycations. We report here that DnaK specifically binds to the polycation protamine when [protamine]/[DnaK] is near unity, whereas protamine induces the aggregations of DnaK when [protamine]/[DnaK] ≥ 20. Complexes between DnaK and protamine were detected using fluorescently labeled protamine (protamine*) in conjunction with high performance size exclusion chromatography. We found that: (i) an unlabeled peptide of known affinity for DnaK partially inhibited the formation of DnaK-protamine* complexes; (ii) Mg-ATP (and Mg-γ-S-ATP) significantly reduced the affinity of protamine* for DnaK; and (iii) the rate of DnaK-protamine* complex dissociation is highly temperature-sensitive, with apparent activation enthalpies (ΔH*) equal to 32 ± 4 and 28 ± 1 kcal mol−1 in the absence of added nucleotide and in the presence of ADP, respectively. The results are consistent with the specific binding of protamine* at the (poly)peptide binding site of DnaK. A model is proposed to account for the protamine-induced aggregation of DnaK.  相似文献   

15.
Monoclonal antibodies specific for the cyclobutane pyrimidine dimer (CPD) are widely used for detection and quantification of DNA photolesions. However, the mechanisms of antigen binding by anti-CPD antibodies are little understood. Here we report NMR analyses of antigen recognition by TDM-2, which is a mouse monoclonal antibody specific for the cis-syn-cyclobutane thymine dimer (T[c,s]T). 31P NMR and surface plasmon resonance data indicated that the epitope recognized by TDM-2 comprises hexadeoxynucleotides centered on the CPD. Chemical shift perturbations observed for TDM-2 Fab upon binding to d(T[c,s]T) and d(TAT[c,s]TAT) were examined in order to identify the binding sites for these antigen analogs. It was revealed that d(T[c,s]T) binds to the central part of the antibody-combining site, while the CPD-flanking nucleotides bind to the positively charged area of the VH domain via electrostatic interactions. By applying a novel NMR method utilizing a pair of spin-labeled DNA analogs, the orientation of DNA with respect to the antigen-binding site was determined: CPD-containing oligonucleotides bind to TDM-2 in a crooked form, draping the 3′-side of the nucleotides onto the H1 and H3 segments, with the 5′-side on the H2 and L3 segments. These data provide valuable information for antibody engineering of TDM-2.  相似文献   

16.
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   

17.
The binding of the novel cytotoxic acridine derivative, 1-[2-(acridin-9-ylamino)ethyl]-1,3-dimethylthiourea (ACRAMTU) to various self-complementary oligonucleotide duplexes has been studied by combined high-resolution NMR spectroscopy/restrained molecular dynamics and equilibrium binding assays to establish the sequence and groove specificity of intercalation. The binding mode in the sequences d(GGACGTCC)2 and d(GGAGCTCC)2 was deduced from chemical shift changes and intermolecular NOEs between the ligand and the oligonucleotides. ACRAMTU intercalated into the 5′-CG/CG and 5′-GA/TC base steps, and penetration of the duplexes occurred from the minor groove. Intercalation of ACRAMTU in d(GGTACC)2 occurs at the central TA/TA step, based on the absence of the internucleotide A4H8–T3H1′ and A4H8–T3H3′ cross-peaks in the 1:1 complex of this sequence. An energy- minimized AMBER model of the 1:2 complex, [d(GGAGCTCC)2(ACRAMTU)2], was generated, which was based on restricted molecular dynamics/ mechanics calculations using 108 NOE distance restraints (including 11 DNA–drug distances per ligand). Equilibrium dialysis experiments were performed using octamers containing various base steps present in the ‘NMR sequences’. The highest affinity for ACRAMTU was observed in d(TATAT ATA)2, followed by d(CGCGCGCG)2 and d(GAG ATCTC)2. The binding levels for CG/CG and GA/TC were virtually the same. The unusual tolerance of the GA/TC intercalation site and the pronounced groove specificity of ACRAMTU play a significant role in the molecular recognition between the corresponding platinum conjugate, Pt-ACRAMTU, and DNA.  相似文献   

18.
Cytosolic extracts of Lactobacillus johnsonii 100-100 (previously reported as Lactobacillus sp. strain 100-100) contain four heterotrimeric isozymes composed of two peptides, α and β, with conjugated bile salt hydrolase (BSH) activity. We now report cloning, from the genome of strain 100-100, a 2,977-bp DNA segment that expresses BSH activity in Escherichia coli. The sequencing of this segment showed that it contained one complete and two partial open reading frames (ORFs). The 3′ partial ORF (927 nucleotides) was predicted by BLAST and confirmed with 5′ and 3′ deletions to be a BSH gene. Thermal asymmetric interlaced PCR was used to extend and complete the 948-nucleotide sequence of the BSH gene 3′ of the cloned segment. The predicted amino acid sequence of the 5′ partial ORF (651 nucleotides) was about 80% similar to the C-terminal half of the largest, complete ORF (1,353 nucleotides), and these two putative proteins were similar to several amine, multidrug resistance, and sugar transport proteins of the major facilitator superfamily. E. coli DH5α cells transformed with a construct containing these ORFs, in concert with an extracellular factor produced by strain 100-100, demonstrated levels of uptake of [14C]taurocholic acid that were increased as much as threefold over control levels. [14C]Cholic acid was taken up in similar amounts by strain DH5α pSportI (control) and DH5α p2000 (transport clones). These findings support a hypothesis that the ORFs are conjugated bile salt transport genes which may be arranged in an operon with BSH genes.  相似文献   

19.
Addition of membrane-permeable cyclic GMP (cGMP) and cyclic AMP (cAMP) were shown to cause elevation of cytosolic Ca2+ concentration ([Ca2+]cyt) in tobacco (Nicotiana plumbaginofolia) protoplasts. Under the same conditions these cyclic nucleotides were shown to provoke a physiological swelling response in the protoplasts. Nonmembrane-permeable cAMP and cGMP were unable to trigger a detectable [Ca2+]cyt response. Cyclic-nucleotide-mediated elevations in [Ca2+]cyt involved both internal and external Ca2+ stores. Both cAMP- and cGMP-mediated [Ca2+]cyt elevations could be inhibited by the Ca2+-channel blocker verapamil. Addition of inhibitors of phosphodiesterases (isobutylmethylxanthine and zaprinast) and the adenylate cyclase agonist forskolin to the protoplasts (predicted to elevate in vivo cyclic-nucleotide concentrations) caused elevations in [Ca2+]cyt. Addition of the adenylate cyclase inhibitor 2′,5′-dideoxyadenosine before forskolin significantly inhibited the forskolin-induced [Ca2+]cyt elevation. Taken together, these data suggest that a potential communication point for cross-talk between signal transduction pathways using cyclic nucleotides in plants is at the level of Ca2+ signaling.  相似文献   

20.
The prevailing hypothesis on the biosynthesis of erucic acid in developing seeds is that oleic acid, produced in the plastid, is activated to oleoyl-coenzyme A (CoA) for malonyl-CoA-dependent elongation to erucic acid in the cytosol. Several in vivo-labeling experiments designed to probe and extend this hypothesis are reported here. To examine whether newly synthesized oleic acid is directly elongated to erucic acid in developing seeds of Brassica rapa L., embryos were labeled with [14C]acetate, and the ratio of radioactivity of carbon atoms C-5 to C-22 (de novo fatty acid synthesis portion) to carbon atoms C-1 to C-4 (elongated portion) of erucic acid was monitored with time. If newly synthesized 18:1 (oleate) immediately becomes a substrate for elongation to erucic acid, this ratio would be expected to remain constant with incubation time. However, if erucic acid is produced from a pool of preexisting oleic acid, the ratio of 14C in the 4 elongation carbons to 14C in the methyl-terminal 18 carbons would be expected to decrease with time. This labeling ratio decreased with time and, therefore, suggests the existence of an intermediate pool of 18:1, which contributes at least part of the oleoyl precursor for the production of erucic acid. The addition of 2-[{3-chloro-5-(trifluromethyl)-2-pyridinyl}oxyphenoxy] propanoic acid, which inhibits the homodimeric acetyl-CoA carboxylase, severely inhibited the synthesis of [14C]erucic acid, indicating that essentially all malonyl-CoA for elongation of 18:1 to erucate was produced by homodimeric acetyl-CoA carboxylase. Both light and 2-[{3-chloro-5-(trifluromethyl)-2-pyridinyl}oxyphenoxy]-propanoic acid increased the accumulation of [14C]18:1 and the parallel accumulation of [14C]phosphatidylcholine. Taken together, these results show an additional level of complexity in the biosynthesis of erucic acid.  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