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.
T G Burke  T R Tritton 《Biochemistry》1985,24(21):5972-5980
We have exploited the intrinsic fluorescence properties of the anthracycline antitumor antibiotics to study the dependence on drug structure of relative drug location and dynamics when the anthracyclines were bound to sonicated dimyristoylphosphatidylcholine (DMPC) and dipalmitoylphosphatidylcholine (DPPC) vesicles at 27.5 degrees C. Iodide quenching experiments at constant ionic strength were used to evaluate the relative accessibilities of the bound fluorophores to membrane-impermeable iodide. Iodide was found to quench the fluorescence of anthracyclines in free solution by both static and dynamic mechanisms, whereas quenching of membrane-bound fluorophores was predominantly due to the dynamic mechanism. Modified Stern-Volmer plots of anthracyclines bound to fluid-phase DMPC bilayers were linear, and the biomolecular rate constant (kq) values ranged from 0.6 X 10(9) to 1.3 X 10(9) M-1 s-1. Modified Stern-Volmer plots of anthracyclines bound to solid-phase DPPC bilayers were curved, indicative of a heterogeneous-bound drug population. A strong correlation between drug hydrophobicity and penetration of the fluorophore into the bilayer was observed for the daunosamine-containing anthracyclines. Steady-state fluorescence anisotropy measurements under iodide quenching conditions were used to investigate the diffusive motions of anthracyclines in isotropic solvent and in fluid-phase DMPC bilayers. Anthracycline derivatives free in solution exhibited limiting anisotropy (alpha infinity) values which decayed to zero at times long compared to the excited-state lifetime, in contrast to anthracyclines bound to fluid-phase DMPC bilayers, which showed nonzero alpha infinity values. Steady-state anisotropies of membrane-bound anthracyclines were found to be governed principally by alpha infinity and not by the mean rotational rate (R).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

2.
The quenching of fluorescence of 5-methoxyindole, N-acetyl-L-tryptophanamide and two single tryptophan containing peptides, melittin and mastoparan X, by oxidized dithiothreitol was studied. The slopes of the Stern-Volmer plots for steady-state fluorescence quenching were 133 M-1, 71.2 M-1, 75.5 M-1 and 35.0 M-1 at 21 degrees C and pH 7.0 for 5-methyoxyindole, N-acetyl-L-tryptophanamide, melittin and mastoparan X respectively. Fluorescence lifetimes of indole or tryptophan in these compounds, as determined by multifrequency phase fluorometry, were decreased by 15% or less at concentrations that produced 50% or more quenching of steady-state fluorescence. Thus, quenching of fluorescence by oxidized dithiothreitol for these derivatives of indole appears to be largely static in nature, suggesting a ground-state interaction.  相似文献   

3.
M R Eftink  D M Jameson 《Biochemistry》1982,21(18):4443-4449
The fluorescence lifetime of liver alcohol dehydrogenase (LADH) has been determined by phase fluorometry at various emission wavelengths and as a function of the concentration of the quencher acrylamide. Acrylamide selectively quenches the fluorescence of the surface tryptophanyl residue Trp-15, thus allowing the fluorescence lifetime of this residue and the buried residue Trp-314 to be evaluated. Values of tau15 = 6.9 ns and tau314 = 3.6 ns are obtained, in qualitative agreement with lifetimes of these residues determined from fluorescence decay studies [Ross, J.B.A., Schmidt, C.J., & Brand, L. (1981) Biochemistry 20, 4369-4377]. The quenching of the fluorescence of LADH by oxygen has also been studied. Quenching by oxygen results in a blue shift in the fluorescence of the protein and a downward-curving Stern-Volmer plot. These data, along with oxygen quenching studies in the presence of 1 M acrylamide, are consistent with a model in which oxygen quenches the fluorescence of Trp-314 and -15 with quenching constants of 3.5 and 25 M-1, respectively. Thus, as in studies with other quenchers, Trp-314 is found to be less accessible to the quencher oxygen than is Trp-15. A lifetime Stern-Volmer plot has also been obtained for the oxygen quenching of LADH. Such a plot deviates somewhat from the intensity Stern-Volmer plot as predicted by simulations of the quenching of two-component systems.  相似文献   

4.
Trichloroethanol is an efficient quencher of indole fluorescence of model compounds and proteins [Eftink, M. R. and Ghiron, C. A. (1976) J. Phys. Chem. 80, 486--493]. At low quencher concentrations, the quenching follows the classical Stern-Volmer law. Bimolecular rate constants calculated from measured quenching constants and lifetimes are equal to 6 X 10(9) M-1s-1 and 1.2 X 10(9) M-1s-1 for N-acetyltrypotophanamide and wheat germ agglutinin, respectively. Upon ultraviolet irradiation in the presence of trichloroethanol, transformation of fluorescent tryptophan occurs, leading to a fluorescent photoproduct. This can be easily used as a method for the quantitative determination of fluorescent tryptophan residues in proteins. In good agreement with previous results, two fluorescent tryptophan residues per polypeptide chain are found in wheat germ agglutinin. Concomitantly with the photochemical reactions, the hemagglutinating protein activity and its affinity constant towards chitin oligomers are reduced. A probable location of tryptophan residues in the binding sites of wheat germ agglutinin is proposed.  相似文献   

5.
A new class of fluorescent nucleotide analogs which contain the fluorophore 1-aminonaphthalene-5-sulfonate attached via a gamma-phosphoamidate bond has been synthesized. Both the purine and pyrimidine analogs have fluorescence emission maxima at 460 nm. Cleavage of the alpha-beta-phosphoryl bond produces change in both the absorption and fluorescence emission spectra. The fluorescence of the pyrimidine analogs is quenched; cleavage of the alpha-beta-phosphoryl bond of the UTP analog produces about a 14-fold increase in fluorescence intensity at 500 nm. Under the same conditions the fluorescence of the CTP analog increases about 8-fold, whereas the fluorescence of the purine analogs shows only a slight change. These derivatives are good substrates for Escherichia coli RNA polymerase with only slightly increased Km values and with Vmax values about 50 to 70% that of the normal nucleotides. They are used less efficiently by wheat germ RNA polymerase II. The ATP analog can be used by E. coli RNA polymerase to initiate RNA chains.  相似文献   

6.
Tryptophan (Trp) fluorescence quenching of phytochrome has been studied using anionic, cationic and neutral quenchers, I-, Cs+ and acrylamide, respectively, in an effort to understand the molecular differences between the Pr and Pfr forms. The data have been analyzed using both Stern-Volmer and modified Stern-Volmer kinetic treatments. The anionic quencher, I-, was proven to be an ineffective quencher with Stern-Volmer constants, Ksv, of 0.60 and 0.63 M-1, respectively, for the Pr and Pfr forms of phytochrome. The cationic quencher, Cs+, showed about a 2-fold difference in the Ksv of Pr and Pfr, indicating a significant change in the fluorescent Trp environments during the Pr to Pfr phototransformation. However, only 25-37% of the fluorescent Trp residues were accessible to the cationic quencher. Most of the fluorescent Trp residues were accessible to acrylamide, but the quenching by acrylamide was indistinguishable for the Pr and Pfr forms. An additional quenching by acrylamide after a saturated quenching with Cs+ showed more than 40% increase in the Ksv of Pfr over Pr. These observations, along with the finding of two distinct components in the Trp fluorescence lifetime, indicate the existence of Trp residues in at least two different sets of environments in the phytochrome protein. The two components of the fluorescence had lifetimes of 1.1 ns (major) and 4.7 ns (minor) for Pr and 0.9 ns (major) and 4.6 ns (minor) for Pfr. Fluorescence quenching was found to be both static and dynamic as the Stern-Volmer constants for the steady-state fluorescence quenching were higher than for the dynamic fluorescence quenching. Based on the quenching results, in combination with the location of Trp residues in the primary structure, we conclude that the Pr to Pfr phototransformation involves a significant conformation change in the phytochrome molecule, preferentially in the 74 kDa chromophore-bearing domain.  相似文献   

7.
Liu R  Siemiarczuk A  Sharom FJ 《Biochemistry》2000,39(48):14927-14938
P-glycoprotein is a member of the ATP binding cassette family of membrane proteins, and acts as an ATP-driven efflux pump for a diverse group of hydrophobic drugs, natural products, and peptides. The side chains of aromatic amino acids have been proposed to play an important role in recognition and binding of substrates by P-glycoprotein. Steady-state and lifetime fluorescence techniques were used to probe the environment of the 11 tryptophan residues within purified functional P-glycoprotein, and their response to binding of nucleotides and substrates. The emission spectrum of P-glycoprotein indicated that these residues are present in a relatively nonpolar environment, and time-resolved experiments showed the existence of at least two lifetimes. Quenching studies with acrylamide and iodide indicated that those tryptophan residues predominantly contributing to fluorescence emission are buried within the protein structure. Only small differences in Stern-Volmer quenching constants were noted on binding of nucleotides and drugs, arguing against large changes in tryptophan accessibility following substrate binding. P-glycoprotein fluorescence was highly quenched on binding of fluorescent nucleotides, and moderately quenched by ATP, ADP, and AMP-PNP, suggesting that the site for nucleotide binding is located relatively close to tryptophan residues. Drugs, modulators, hydrophobic peptides, and nucleotides quenched the fluorescence of P-glycoprotein in a saturable fashion, allowing estimation of dissociation constants. Many compounds exhibited biphasic quenching, suggesting the existence of multiple drug binding sites. The quenching observed for many substrates was attributable largely to resonance energy transfer, indicating that these compounds may be located close to tryptophan residues within, or adjacent to, the membrane-bound domains. Thus, the regions of P-glycoprotein involved in nucleotide and drug binding appear to be packed together compactly, which would facilitate coupling of ATP hydrolysis to drug transport.  相似文献   

8.
B Turk  R Awad  E V Usova  I Bj?rk  S Eriksson 《Biochemistry》1999,38(26):8555-8561
Deoxycytidine kinase (dCK) is an enzyme with broad substrate specificity which can phosphorylate pyrimidine and purine deoxynucleosides, including important antiviral and cytostatic agents. In this study, stopped-flow experiments were used to monitor intrinsic fluorescence changes induced upon binding of various phosphate donors (ATP, UTP, and the nonhydrolyzable analogue AMP-PNP) and the acceptor dCyd to recombinant dCK. Monophasic kinetics were observed throughout. The nucleotides as well as dCyd bound to the enzyme by a two-step mechanism, involving a rapid initial equilibrium step, followed by a protein conformational change that is responsible for the fluorescence change. The bimolecular association rate constants for nucleotide binding [(4-10) x 10(3) M-1 s-1] were 2-3 orders of magnitude lower than those for dCyd binding [(1.3-1.5 x 10(6) M-1 s-1]. This difference most likely is due predominantly to the large difference in the forward rate constants of the conformational changes (0.04-0.26 s-1 vs 560-710 s-1). Whereas the kinetics of the binding of ATP, UTP, and AMP-PNP to dCK showed some differences, UTP exhibiting the tightest binding, no significant differences were observed for the binding of dCyd to dCK in the presence or absence of phosphate donors. However, the binding of dCyd to dCK in the presence of ATP or UTP was accompanied by a 1.5- or 3-fold higher quenching amplitude as compared with dCyd alone or in the presence of AMP-PNP. We conclude that ATP and UTP induce a conformational change in the enzyme, thereby enabling efficient phosphoryl transfer.  相似文献   

9.
The presence of adenine nucleotide translocase (ANT) was found to greatly enhance the partitioning of the ATP analog 2',3'-O-(2,4,6-trinitrophenyl)-adenosine 5'-triphosphate (TNP-ATP) into reduced Triton X-100 micelles. The protein's effect was studied through the quenching of fluorescence of purified ANT, irreversibly inhibited by carboxyatractyloside (CAT), solubilized in reduced Triton X-100 micelles. The dependence of quenching of the protein's time-resolved tryptophan fluorescence on TNP-ATP concentration was measured and found to follow a Stern-Volmer mechanism. However, the calculated quenching constant was too large to be accounted for by the aqueous TNP-ATP concentration. Experiments were therefore conducted to determine the partitioning of the quencher between the three phases present: aqueous, protein-free micelle, and protein micelle; a system also described by the equation of Omann, G. M., and M. Glaser (1985. Biophys. J. 47:623-627.). By measuring the dependence of the apparent quenching rate constant on the protein concentration and protein/micelle ratios, this equation was used to calculate both the quencher partition coefficient into protein-free micelles (Pm) and into protein-micelles (Ppm), as well as the bimolecular quenching rate constant (kpm) in protein micelles. From the quenching experiments, kpm = 5.0 x 10(8)M-1s-1,Pm = 290 and pyrene quenching experiment to be 325, and by a rapid filtration experiment to be 450. Clearly, the presence of the integral membrane protein ANT-CAT in reduced Triton X-100 micelles greatly increases the partition of TNP-ATP into the micelle. ANT alters the properties and thus, the structure of the detergent micelle, which has direct implications for the use of detergent micelles as a model system for membrane proteins and may indicate that analogous effects occur in the mitochondrial membrane.  相似文献   

10.
Since de‐novo synthesis of pyrimidine nucleotides is coupled to the mitochondrial respiratory chain (RC) via dehydroorotic acid dehydrogenase (DHODH), respiratory chain dysfunction should impair pyrimidine synthesis. To investigate this, we used specific RC inhibitors, Antimycin A and Rotenone, to treat primary human keratinocytes and 143B cells, a human osteosarcoma cell line, in culture. This resulted in severe impairment of de novo pyrimidine nucleotide synthesis. The effects of RC inhibition were not restricted to pyrimidine synthesis, but concerned purine nucleotides, too. While the total amount of purine nucleotides was not diminished, they were significantly broken down from triphosphates to monophosphates, reflecting impaired mitochondrial ATP regeneration. The effect of Rotenone was similar to that of Antimycin A. This was surprising since Rotenone inhibits complex I of the respiratory chain, which is upstream of ubiquinone where DHODH interacts with the RC. In order to avoid unspecific effects of Rotenone, we examined the consequences of a mitochondrial DNA mutation that causes a specific complex I defect. The effect was much less pronounced than with Rotenone, suggesting that complex I inhibiton cannot fully explain the marked effect of Rotenone on pyrimidine nucleotide synthesis.  相似文献   

11.
Since de-novo synthesis of pyrimidine nucleotides is coupled to the mitochondrial respiratory chain (RC) via dehydroorotic acid dehydrogenase (DHODH), respiratory chain dysfunction should impair pyrimidine synthesis. To investigate this, we used specific RC inhibitors, Antimycin A and Rotenone, to treat primary human keratinocytes and 143B cells, a human osteosarcoma cell line, in culture. This resulted in severe impairment of de novo pyrimidine nucleotide synthesis. The effects of RC inhibition were not restricted to pyrimidine synthesis, but concerned purine nucleotides, too. While the total amount of purine nucleotides was not diminished, they were significantly broken down from triphosphates to monophosphates, reflecting impaired mitochondrial ATP regeneration. The effect of Rotenone was similar to that of Antimycin A. This was surprising since Rotenone inhibits complex I of the respiratory chain, which is upstream of ubiquinone where DHODH interacts with the RC. In order to avoid unspecific effects of Rotenone, we examined the consequences of a mitochondrial DNA mutation that causes a specific complex I defect. The effect was much less pronounced than with Rotenone, suggesting that complex I inhibiton cannot fully explain the marked effect of Rotenone on pyrimidine nucleotide synthesis.  相似文献   

12.
Daunomycin and adriamycin were previously found to produce Q-like banding patterns on chromosomes. The interaction of several anthracyclines with both natural and synthetic DNAs and chromosomes has been investigated in more detail. Daunomycin fluorescence is almost completely quenched by natural DNAs with varying base composition from 31 to 72% G-C and by the alternating polymer poly-d(G-C)·poly-d(G-C). In contrast, daunomycin fluorescence is quenched by only 50% when the dye interacts with synthetic A-T polymers. Thus, differential quenching of daunomycin fluorescence can account for the production of bright bands at contiguous A-T sequences along the chromosome. Slight differences in fluorescence quenching between the repeating and homopolymeric A-T duplex DNAs were observed which can be attributed to differences in affinity of daunomycin for these DNAs. The aminosugar moiety of daunomycin, daunosamine, increases the binding of daunomycin to DNA and also enhances chromosome banding. — Nogalamycin, which displays no differential quenching with the different DNAs in solution, also fails to produce bands on chromosomes. — These findings suggest that non-random nucleotide sequence arrangements along the chromosome are a basic determinant for dye interaction to produce the observed banding patterns. Specific banding procedures may determine the accessibility of these sites within the chromosomal DNA.  相似文献   

13.
The binding of daunomycin and its Bolton-Hunter derivative iodomycin to plasma membranes isolated from multidrug-resistant Chinese hamster ovary cells (CHO B30) and their drug-sensitive parents (B1) was investigated. The thermodynamics and kinetics of equilibrium binding monitored by fluorescence titrations and temperature-jump relaxation spectrometry were compared with the specificity of covalent photolabeling with [3H]daunomycin and [125I]iodomycin. The facts that the uptake of anthracycline from aqueous solution into the CHO membranes was not accompanied by any substantial increase of fluorescence anisotropy nor by any spectral shift of the fluorescence emission spectrum and that the partition ratio into the membrane was 20-30-fold higher when compared to a lecithin bilayer, provided evidence that the non-covalent drug binding sites are constituted by polar protein domains without any substantial contribution from the surrounding lipids. Photoaffinity labeling with nanomolar concentrations of anthracycline and equilibrium binding curves independently showed that a 150-170-kDa plasma membrane glycoprotein (P-glycoprotein), whose overexpression is the major difference between B1 and B30 membranes, provides the binding sites of highest affinity for daunomycin and iodomycin (K approximately equal to 4 x 10(7) M-1). Comparison of photolabeling and equilibrium data suggested that the same binding sites on P-glycoprotein were most probably being monitored. The photolabeling of P-glycoprotein by iodomycin was inhibited in a dose-dependent manner by other compounds to which multi-drug-resistant cells are either resistant or collaterally sensitive with the following orders of effectiveness: vinblastine greater than verapamil greater than nitrendipine greater than daunomycin much greater than colchicine. Temperature-jump experiments covering the time range of 1 microseconds to 1 s revealed a single concentration-dependent relaxation time of 10-30 microseconds. The association of daunomycin with its binding sites in the membranes was found to be a diffusion-controlled process with kon rates of 2-4 X 10(9) M-1 s-1. Therefore, the selectivity of drug binding was entirely reflected in the dissociation rates.  相似文献   

14.
A class of N-substituted quinoline compounds has been introduced recently for the fluorescence measurement of Cl concentration in biological preparations. The most Cl-sensitive compound was 6-methoxy-N-[3-sulfopropyl] quinolinium with peak excitation and emission wavelengths of 350 and 442 nm and a Stern-Volmer constant for quenching by Cl of 118 M-1. Six water-soluble quinoline derivatives were synthesized and characterized for the purposes of increasing Cl sensitivity, adding ester functions for cell trapping, and red-shifting the fluorescence peak wavelengths. Acetic acid ester functions were added at the N-, 2-, and 6-positions of the quinoline ring. The best ester compound, N-(6-methoxyquinolyl)acetoethyl ester (MQAE), was water soluble (270 g/liter at 23 degrees C; octanol:H2O partition coefficient of 0.009), had a high Cl sensitivity (Stern-Volmer constant 200 M-1), peak excitation and emission wavelengths of 355 and 460 nm, a fluorescence lifetime of 21.6 ns, and a molar absorbance of 4850 M-1 cm-1 (320 nm). MQAE fluorescence was not altered by the physiological anions HCO3, SO4, and PO4, by cations, or by pH. MQAE was used to measure chloride transport in liposome membranes and in cultured LLC-PK1 cells in monolayer; MQAE leaked out of cells less than 20% in 60 min at 37 degrees C. The physical, optical, and anion quenching properties for the series of ester compounds were determined to establish a set of structure-activity correlates.  相似文献   

15.
Defects in X-linked phosphoribosylpyrophosphate synthetase 1 (PRPS1) manifest as follows: (1) PRS-I enzyme "superactivity" (gain-of-function mutations affecting allosteric regions); (2) PRS-I overexpression (which may be linked to miRNA mutation); (3) severe PRS-I deficiency/Arts syndrome (missense mutations producing loss-of-function); (4) moderate PRS-I deficiency/Charcot-Marie-Tooth disease-5 (less severe loss-of-function mutations); and (5) mild PRS-I deficiency/Deafness-2 (mutations producing slight destabilization). Similar to Lesch-Nyhan disease, PRPS1-related disorders arise from phosphoribosyl-pyrophosphate (PRPP)-dependent nucleotide "depletion" of purine nucleotides (e.g., ATP, GTP). S-adenosylmethionine (SAMe) appears to partially alleviate purine depletion via a PRPP-independent path. Synthesis of pyrimidine nucleotides is PRPP dependent, with uridine monophosphate synthase deficiency producing pyrimidine nucleotide depletion. But pyrimidine salvage from uridine does not require PRPP, and this nucleoside is transported freely to pyrimidine-depleted tissues. Regulation of nicotinamide nucleotides is less clear; synthesis from pyridine nucleobases is PRPP dependent. Nucleotide "depletion" contrasts with nucleotide "toxicity," exemplified by the purine disorders adenosine deaminase (ADA) and purine nucleoside phosphorylase (PNP) deficiencies or by pyrimidine nucleotidase deficiency. These are characterized by the accumulation of one or more abnormal nucleotides such as succinyl- or deoxy-nucleotides or their metabolites, which interrupt other nucleotide or related pathways or are toxic to specific cell types. Theoretically, purine toxicity disorders would not be ameliorated by SAMe therapy, and this was confirmed for one adenylosuccinate lyase-deficient child. Nucleotide defects may also be seen as an aspect of mitochondrial disease, with SAMe-based mitochondrial therapy perhaps meriting further investigation.  相似文献   

16.
M Lamkin  T Tao  S S Lehrer 《Biochemistry》1983,22(13):3053-3058
Rabbit skeletal alpha alpha-tropomyosin was specifically labeled at Cys-190 with the fluorescent probe N-(iodoacetyl)-N'-(1-naphthyl-5-sulfo)ethylenediamine (1,5-IAE-DANS). The fluorescence decay of the resultant AE-DANS-labeled alpha alpha-tropomyosin (Tm) was monoexponential with a lifetime of 13.55 ns. When acrylamide was used as the quencher, the apparent Stern-Volmer quenching constant Ksv' for Tm was measured to be 5.78 M-1 and the quenching rate constant kq to be 3.20 X 10(8) M-1 s-1. The presence of troponin reduced the magnitude of Ksv' to 4.14 M-1 and induced the appearance of a second decay component. This second component had an amplitude of approximately 20% of the total intensity, a lifetime of approximately 20 ns, and a kq of 4.5 X 10(-7) M-1 s-1. Similarly, the presence of F-actin induced the appearance of a minor longer lived decay component with a decreased kq. On the basis of the increase in the lifetime and the decrease in kq, the appearance of the long-lived decay component was interpreted to be due to troponin or actin interacting with Tm near the Cys-190 site in both cases. Our results further suggest that the label was capable of equilibrating between an exposed hydrophilic environment on the surface of Tm and a buried hydrophobic environment at the troponin-Tm or actin-Tm interaction interfaces.  相似文献   

17.
18.
19.
The intrinsic fluorescence of lauryl maltoside solubilized bovine heart cytochrome c oxidase has been determined to arise from tryptophan residues of the oxidase complex. The magnitude of the fluorescence is approximately 34% of that from n-acetyltryptophanamide (NATA). This level of fluorescence is consistent with an average heme to tryptophan distance of 30 A. The majority of the fluorescent tryptophan residues are in a hydrophobic environment as indicated by the fluorescence emission maximum at 328 nm and the differing effectiveness of the quenching agents: Cs+, I-, and acrylamide. Cesium was ineffective up to a concentration of 0.7 M, whereas quenching by the other surface quenching agent, iodide, was complex. Below 0.2 M, KI was ineffective whereas between 0.2 and 0.7 M 15% of the tryptophan fluorescence was found to be accessible to iodide. This pattern indicates that protein structural changes were induced by iodide and may be related to the chaotropic character of KI. Acrylamide was moderately effective as a quenching agent of the oxidase fluorescence with a Stern-Volmer constant of 2 M-1 compared with acrylamide quenching of NATA and the water-soluble enzyme aldolase having Stern-Volmer constants of 12 M-1 and 0.3 M-1, respectively. There was no effect of cytochrome c on the tryptophan emission intensity from cytochrome c oxidase under conditions where the two proteins form a tight, 1:1 complex, implying that the tryptophan residues near the cytochrome c binding site are already quenched by energy transfer to the homes of the oxidase. The lauryl maltoside concentration used to solubilize the enzyme did not affect the fluorescence of NATA.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
The thermodynamic parameters of the interaction of the two anthracyclines 13-dihydrodaunomycin and marcellomycin with calf thymus DNA were examined by equilibrium binding studies. Enthalpy and entropy changes of the binding of both drugs show salt dependence profiles that cannot be rationalized by the polyelectrolyte theory. This feature is common to other anthracycline compounds. The nucleotide sequence binding preferences of daunomycin, adriamycin, 13-dihydrodaunomycin and marcellomycin have been studied by monitoring the degree of protection from cleavage by restriction endonucleases of linearized pBR322. Differential protection of pBR322 DNA against the cleavage of Bgl I and Ava II suggests that these drugs recognize changes in the sequences near the enzyme recognition site. Alterations of the electrophoretic restriction pattern of pBR322 in the presence of anthracyclines are dependent on time and on concentration. These results are discussed in relation to the existence of nucleotide sequences with different affinity for these drugs.  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