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.
Metabolic inhibitors were applied after the transport system was fully developed in concentrations sufficient to block cleavage. 0.5–1.0 × 10?4 M cyanide and anaerobiosis caused from negligible to moderate (40%) inhibition of phosphate uptake. The inhibition occurred late in the breeding season, and the inhibitory action of cyanide on uptake was associated with irreversible developmental effects. Azide (3 × 10?3 M) did not inhibit uptake when the chamber method was used, but the aliquot and Hopkins' tube methods gave considerable inhibition. Purified preparations of 2,4-dinitrophenol (1 × 10?4 M) did not inhibit uptake. Sodium iodoacetate (up to 0.05 M) and phlorizin (0.005 M) exerted no effect. Calculations of the minimal work requirement for the transport process reveal that this amounts to only a small fraction (0.24% at an external phosphate concentration of 2 μM) of the total available metabolic energy. Exposure of eggs at five minutes after insemination (lag phase) to cyanide (5 × 10?5 M), anaerobic conditions, or azide (3 × 10?3 M) blocked the expected increase of phosphate uptake. Removal of the inhibitors led to resumption of development and the appearance of the phosphate transport system in an essentially normal pattern. Exposure of eggs to 1.4–2.0 × 10?4 M p-chloromercuribenzoate (p-CMB) during the accumulation phase severely depressed phosphate uptake, but cleavage was not inhibited nor delayed; recovery from the inhibition was accelerated by 1 × 10?3 M cysteine. Exposure to p-CMB during the lag phase blocked the appearance of the transport system; cleavage proceeded normally. After the removal of p-CMB little reversal occurred until the addtion of 1 × 10?3 M cysteine, when the phosphate transport system developed in an essentially normal manner. Trypsin (0.001–0.01%) neither activates the transport system in unfertilized eggs, nor inactivates it in denuded fertilized eggs by removal of surface proteins. The data are consistent with the conclusion that (1) the phosphate transport system is newly synthesized at fertilization in energy dependent reactions, and (2) phosphate transport is a carrier mediated process not directly dependent on metabolic energy.  相似文献   

2.
The amino-reactive reagent, 4-acetamido-4'-isothiocyanostilbene-2,2'-disulfonic acid (SITS),1 considerably reduces the uptake of the sulfhydryl agent, parachloromercuriphenylsulfonic acid (PCMBS), but does not reduce its effects on cation permeability and on cation transport. These data indicate that PCMBS enters the membrane by at least two channels, one sensitive and the other insensitive to SITS, with only the latter leading to the cation-controlling sulfhydryl groups. Substitution of phosphate or sulfate for chloride results in an inhibition of PCMBS uptake via the SITS-insensitive pathway. These and other data lead to the conclusion that the SITS-sensitive pathway is the predominant one for anion permeation, and the insensitive one for cation permeation. Parachloromercuribenzoate (PCMB), an agent that is more lipid-soluble than PCMBS, penetrates faster but has a smaller effect on cation permeability. Its uptake is less sensitive to SITS. These and other observations suggest that the cation permeation path involves an aqueous channel in the membrane.  相似文献   

3.
We have investigated the effects of the amino reactive reagent, 2,4,6-trinitrobenzene sulfonic acid (TNBS) on anion transport (chloride and sulfate) and on the K+ content of Ehrlich ascites tumor cells. Incubation of tumor cells with TNBS (3 mM or 10 mM) results in a time dependent uptake of this molecule. Tightly bound TNBS caused a loss of K+ as well as inhibition of sulfate uptake. Although sulfate transport was inhibited by tightly bound TNBS (40% inhibition with 20 nmoles bound per 107 cells), reversibly bound TNBS exerted much greater inhibition. Kinetic analysis of sulfate transport in the presence and absence of TNBS suggests that: (1) tightly bound TNBS exerts a competitive inhibition by occupying membrane sites remote from the specific transport site, (2) TNBS reversibly interacts with a separate site also in a competitive fashion. Increasing amounts of tightly bound TNBS resulted in an enhanced chloride influx. However, reversibly bound TNBS was without effect. These results are in contrast to the effect of TNBS on sulfate transport and show that TNBS, at least in this cell type, is not a general inhibitor of anion transport.  相似文献   

4.
Properties of the fully developed phosphate transport system in the fertilized egg of the sea urchin, Strongylocentrotus purpuratus, were investigated. The rates of phosphate transport at concentrations of external phosphate of 1 to 44 μM, both in the absence and in the presence of 100 μM arsenate, exhibit typical saturation kinetics. At sea water concentrations of 2 μM phosphate, the rate of uptake is about 2 × 10?9 μm/egg/minute at 15°C. Arsenate is a competitive inhibitor of phosphate transport, fully and immediately reversible in its effects, yielding Ki values ranging from 10.5 to 14.1 × 10?6 M in comparison to the corresponding apparent KM (Michaelis-Menten) constants for phosphate of 5.6 to 7.5 × 10?6 M (pH 8.0, 15°C). The rate of arsenate uptake in a phosphate deficient medium amounts to 2.8 to 2.9 × 10?10 μm arsenate/egg/minute at an arsenate concentration of 2.9 to 10.2 μM arsenate (HAsO4??), which is 9.5 and 5.6% of the rate of phosphate uptake at corresponding phosphate concentrations. Arsenate has essentially the same developmental effects at initial concentrations of 5–10 μM and 100 μM arsenate, namely no observable effects for exposure periods of 7.5 hours, although longer periods result in blockage of development at the early blastula stage. Outward flux of phosphate ions cannot be demonstrated by washing prelabelled eggs with sea water containing low or high concentrations of phosphate, even when phosphorylation has been blocked by exposing the eggs to a metabolic inhibitor. Phosphate uptake rates measured in the pH range from 5.0 to 10.0 reveal a sharp optimum at pH 8.8–8.9. Reference to the apparent pK' values of the phosphoric acid system indicate that the entering species is the HPO4?? ion. The effects on rates of phosphate uptake of exposure to sea water at pH values between 7 and 10 for 30 minute periods are fully reversible, but at lower pH values, reversal is delayed, and is only partial. Sodium molybdate (0.01 M), sodium pyrophosphate (1.5 × 10?4 M), and adenosine triphosphate (1–5 × 10?4 M) for exposure periods ranging from 40 to 180 minutes did not significantly affect phosphate uptake. Omission of Ca++ ion from artificial sea water is without effect on phosphate uptake but the absence of both Ca++ and Mg++ results in profound and irreversible depression of both phosphate uptake and development. The data of this and the following paper are consistent with the conclusion that the transport of phosphate involves a surface located carrier. The apparent secondary and tertiary ionization constants of phosphoric acid in sea water (ionic strength = 0.6885) were measured, resulting in a value for pK′2 = 6.14 and for pK′3 = 10.99, at 15°C and phosphate at infinite dilution.  相似文献   

5.
The permeabilities of sarcoplasmic reticulum vesicle membrane for various ions and neutral molecules were measured by following the change in light scattering intensity due to the osmotic volume change of the vesicles. 4-Acetoamido-4′-isothiocyanostilbene-2,2′-disulfonate (SITS), which is a potent inhibitor for the anion permeability of red blood cells membrane, inhibited the permeability of sarcoplasmic reticulum for anions such as Cl?, Pi and methanesulfonate, while it slightly increased that for cations and neutral molecules such as Na+, K+, choline and glycerol. Binding of 5μmol SITS/g protein was necessary for the inhibition of anion permeability. These results suggest the existence of a similar anion transport system in sarcoplasmic reticulum membrane as revealed in red blood cell membrane.  相似文献   

6.
The inactivation of phosphorylase phosphatase by fluorophosphate is described. The inactivation is dependent upon time and concentration of fluorophosphate and cannot be reversed by removal of fluorophosphate from the enzyme. Acid hydrolysis of fluorophosphate destroys the capacity for inhibition. The inactivation exhibits saturation kinetics. A dissociation constant for the enzyme-fluorophosphate complex and a rate constant for the reaction were calculated to be 5.5 × 10?3 M and 0.22 min?1, respectively. A competitive inhibitor, phosphate, protects the enzyme against inactivation. The data are consistent with an irreversible covalent modification of the active site of phosphorylase phosphatase by fluorophosphate.  相似文献   

7.
The effects of phloretin, H2DIDS (4,4′-diisothiocyano-1,2-diphenylethane-2,2′-disulfonate) and SO4?2 on anion transport in Ehrlich ascites tumor cells was studied in an effort to determine whether Cl? and SO4?2 share a common transport mechanism. Sulfate, in the presence of constant extracellular Cl? (100 mM), reduces Cl? self-exchange by 43% (40 mM SO4?2) and Cl??SO4?2 exchange by 36% (25 mM Cl?/O SO4?2) compared to 25 mM Cl?/50 mM SO4?2. Phloretin blocks without delay and to the same extent the self-exchange of both Cl? and SO4?2. For example, at 10?4 M phloretin, anion transport is inhibited 28% which increases to 78% at 5 × 10?4 M. Reversibly bound H2DIDS also inhibits the self-exchange of both Cl? and SO4?2. However, at all H2DIDS concentrations tested (0.5 ? 10 × 10?5 M) SO4?2 transport was far more susceptible to inhibition than that of Cl?. H2DIDS when irreversibly bound to the cell inhibits SO4?2 but not Cl? transport The results of these experiments are consistent with the postulation that both Cl? and SO4?2 are transported by a common mechanism possessing two reactive sites.  相似文献   

8.
The effect of extracellular Pi and arsenate on Pi-transport in Ehrlich ascites tumor cells has been studied. Pi-transport can be described by Michaelis-Menten kinetics; the maximal flux equal to 44 mmoles (kg cell water)?1 hour?1 and Km equal to 3.3 × 10?4 M . Arsenate is a competitive inhibitor of Pi-transport with an inhibition constant (Ki) equal to 2.41 × 10?3 M . The data support the hypothesis that cellular Pi is regulated by the cell membrane through the mediation of a carrier system.  相似文献   

9.
The effects of furosemide and 4-acetamido-4′-isothiocyanostilbene-2,2′-disulfonic acid (SITS) on steady-state Cl flux were studied in Ehrlich mouse ascites cells. At 10 mM, furosemide inhibited isotopically-determined Cl flux by 86% without changing cell Cl content, indicating that influx and efflux were depressed by the same amount. These results suggest that at least 86% of the steady-state Cl flux may occur as a one for one exchange. Half of the inhibitory effect was not reversed by vigorous washing with albumin-Ringer. A smaller portion of steady-state Cl flux was inhibited by SITS. The maximum effect of SITS was reached near 0.6 mM; at this concentration Cl flux was reduced by 37% without an alteration in cell Cl content. Possible competition of environment Cl and SITS was investigated by replacing environment Cl with acetate or NO3. These anions reduced the efficacy of SITS because they depressed cell Cl turnover themselves, apparently acting on the same exchange process.  相似文献   

10.
The Raman spectra of guanylyl (3′-5′) guanosine (GpG) in solution in H2O and D2O at pH 3–7 have been recorded at various temperatures between 0 and 80°C. The results are consistent with the existence in the lower temperature range of stable aggregates formed by the stacking of GpG tetramers. The aggregates melt cooperatively near 60°C, which results in important changes in the spectra. Among these, a large increase in intensity of some of the bands assigned to the guanine residues shows that unstacking of the bases occurs at the melting. Also apparent in the spectra are changes in the intensity and frequency of band attributable to molecular groups involved in intermolecular hydrogen bonding between adjacent molecules in the complex. The melting temperature of GpG decreases by approximately 15°C upon lowering the concentration from 5 × 10?2 to 5 × 10?4M, as shown by Raman, calorimetric, CD, and uv measurements. The experimentally determined ΔH and ΔS for the melting transition are 9 Kcal/mol and 28 e.u./mol, respectively. The aggregation of GpG in 1.5 × 10?3M solutions was found to be very slow. The half-time of the process, which roughly follows first-order kinetics, is approximately 3 min at 10°C and 21 min at 35°C. The negative energy of activation associated with this reaction (?143 Kcal) indicated that the process involves intermediates whose concentrations decrease the temperatures raised, thus slowing down the overall process. The rate of disaggregation of GpG upon dilution to very low concentration is also extremely slow, indicating that the GpG aggregates, once formed, are very stable.  相似文献   

11.
Quenching effects of bergenin, based on the electrochemiluminescence (ECL) of the tris(2,2′‐bipyridyl)‐ruthenium(II) (Ru(bpy)32+)/tri‐n‐propylamine (TPrA) system in aqueous solution, is been described. The quenching behavior can be observed with a 100‐fold excess of bergenin over Ru(bpy)32+. In the presence of 0.1 m TPrA, the Stern–Volmer constant (KSV) of the ECL quenching is as high as 1.16 × 104 M?1 for bergenin. The logarithmic plot of the inhibited ECL versus logarithmic plot of the concentration of bergenin was linear over the range 3.0 × 10?6–1.0 × 10?4 mol/L. The corresponding limit of detection was 6.0 × 10?7 mol/L for bergenin (S/N = 3). In the mechanism of quenching it is believed that the competition of the active free radicals between Ru(bpy)32+/TPrA and bergenin was the key factor for the ECL inhibition of the system. Photoluminescence, cyclic voltammetry, coupled with bulk electrolysis, supports the supposition mechanism of the Ru(bpy)32+/TPrA–bergenin system. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

12.
An inhibitory effect of potassium ferricyanide on the light-dependent active uptake of chloride anions into the alga Hydrodictyon reticulatum has drawn attention to the possibility that chloride transport proceeds as an exchange for electrons and hence can be impaired by artificial electron acceptors. Tetrathiafulvalene radical cation present in solutions of the highly conductive TTF-C1 crystals is easily reduced to TTF0 and inhibits the chloride uptake more powerfully than does ferricyanide. A practically complete inhibition of Cl? uptake is brought about by 3 × 10?5 M TTF+. The presence of ferricyanide actually reduces the inhibitory effect of TTF+; the complex equilibria of a mixture of the two electron acceptors were examined by cyclic voltammetry and a competition between the two for the inhibitory function has been suggested as a possible explanation.  相似文献   

13.
The kinetics of the inhibition of mouse brain glutamate decarboxylase by pyri-doxaI-5′-phosphate oxime-O-acetic acid (PLPOAA) was studied. The inhibition was noncompetitive with regard to glutamic acid; it could be partially reversed by pyridoxal phosphate, but only when the concentration of the latter in the incubation medium was higher than that of pyridoxal-5′-phosphate oxime-O-acetic acid. The inhibition produced by aminooxyacetic acid, which is remarkably greater than that produced by PLPOAA, was also partially reversed only when an excess of pyridoxal phosphate was added. Both in the presence and in the absence of a saturating concentration of pyridoxal phosphate, the activity of the enzyme was decreased by PLPOAA at a 10?4m concentration to a value of about 50 per cent of the control value obtained without added coenzyme. This activity could not be further reduced even when PLPOAA concentration was increased to 5 × 10?3m . This same minimal activity of glutamate decarboxylase was obtained after dialysis of the enzymic preparation, or after incubation with glutamic acid in the cold followed by filtration through Sephadex G-25. The addition of pyridoxal phosphate to the dialysed or glutamic acid-treated enzyme restored the activity to almost the control values. PLPOAA did not affect the activity of glutamate decarboxylase from E. coli or that of DOPA decarboxylase and GABA transaminase from mouse brain. To account for the results obtained it is postulated that brain glutamate decarboxylase has two types of active site, one with firmly bound, non-dialysable pyridoxal phosphate and the other with loosely bound, dialysable coenzyme; PLPOAA behaves as a weak inhibitor probably because it can combine mainly with the loosely bound coenzyme site, while aminooxyacetic acid is a potent inhibitor probably because it can block both the ‘loosely bound coenzyme’ and the ‘firmly bound coenzyme’ sites.  相似文献   

14.
The Cl? transport properties of the luminal border of bovine tracheal epithelium have been investigated using a highly purified preparation of apical plasma membrane vesicles. Transport of Cl? into an intravesicular space was demonstrated by (1) a linear inverse correlation between Cl? uptake and medium osmolarity and (2) complete release of accumulated Cl? by treatment with detergent. The rate of Cl? uptake was highly temperature-sensitive and was enhanced by exchange diffusion, providing evidence for a carrier-mediated transport mechanism. Transport of Cl? was not affected by the ‘loop’ diuretic bumetanide or by the stilbene-derivative anion-exchange inhibitors SITS (4-acetamido-4′-isothiocyanostilbene-2,2′-disulfonic acid) and DIDS (4,4′-diisothiocyanostilbene-2,2′-disulfonic acid). In the presence of the impermeant cation, tetramethylammonium (TMA+), uptake of Cl? was minimal; transport was stimulated equally by the substitution of either K+ or Na+ for TMA+. Valinomycin in the presence of K+ enhanced further Cl? uptake, while amiloride reduced Na+-stimulated Cl? uptake towards the minimal level observed with TMA+. These results suggest the following conclusions: (1) the tracheal vesicle membrane has a finite permeability to both Na+ and K+; (2) the membrane permeability to the medium counterion determines the rate of Cl? uptake; (3) Cl? transport is not specifically coupled with either Na+ or K+; and, finally (4) Cl? crosses the tracheal luminal membrane via an electrogenic transport mechanism.  相似文献   

15.
Mono-, di-, and trisulfonic acids, including 4,4′-diacetamido stilbene-2,2′-disulfonic acid (DAS) and 2-(4′-amino phenyl)-6-methylbenzene thiazol-3′,7-disulfonic acid (APMB) produce a reversible inhibition of sulfate equilibrium exchange in human red cells. A study of the sidedness of the action of a number of these sulfonic acids in red cell ghosts revealed that some, like DAS, inhibit only at the outer membrane surface while others, like APMB, inhibit at either surface. This finding suggests that at least two different types of membrane sites are involved in the control of anion permeability. The nature of the anion permeability controlling sites in the outer cell surface was investigated by studying the effects of DAS on the inhibition by dinitrofluoro-benzene (DNFB) of anion equilibrium exchange and on the binding of DNFB to the proteins of the red blood cell membrane. After exposure to DNFB in the presence of DAS for a certain period of time, there was a reduction of both the inhibitory effect of DNFB on sulfate exchange and the binding of DNFB to the protein in band 3 of SDS polyacrylamide gel electropherograms (nomenclature of Steck, J. Cell. Biol., 62: 1, 1974). Since binding to other membrane proteins was not affected, this observation supports the assumption that the protein in band 3 plays some role in anion transport. In accordance with the absence of an inhibitory effect at the inner membrane surface, internal DAS does not affect DNFB binding to the protein in band 3. DAS protected the anion exchange system not only against inhibition by DNFB but also by m-isothiocyanato benzene sulfonic acid. In contrast to DAS, the equally inhibitory phlorizin does not reduce the rate of dinitrophenylation of the protein in band 3. This suggests that either not all inhibitors of anion exchange exert their action by a combination with sites on the protein in band 3 or that in spite of the described evidence this protein is not involved in the control of anion movements. The effect of the irreversibly binding inhibitor 4-acetamido-4′-isothiocyanato-stilbene-2,2′-disulfonic acid (SITS) on DNFB binding to the protein in band 3 was studied in an attempt to differentiate DNFB binding related to inhibition of anion permeability from DNFB binding which is not involved. At least three distinguishable populations of DNFB binding sites were found: (1) binding sites common for DNFB and SITS which are probably related to inhibition, (2) other common sites which are not related to inhibition and (3) different sites whose dinitrophenylation is not affected by SITS. The number of sites in population (1) was estimated to be 0.8–1.2 ± 106/cell. A study of the concentration dependence of the inhibition of anion equilibrium exchange with 4,4′-isothiocyanato-2,2′-stilbene disulfonic acid (DIDS) and APMB further suggests that among the sites in population (1) a major fraction is susceptible to modification by APMB and DIDS while the rest is only susceptible to DIDS. It remains undecided whether these differences of susceptibility reflect differences of accessibility or reactivity.  相似文献   

16.
Potato tuber phosphofructokinase was purified 19·.6-fold by a combination of ethanol fractionation and DEAE-cellulose column chromatography. The enzyme was very unstable; its pH optimum was 8·0. Km for fructose-6-phosphate, ATP and Mg2+ was 2·1 × 10?4 M, 4·5 × 10?5 M and 4·0 × 10?4 M respectively. ITP, GTP, UTP and CTP can act as phosphate donors, but are less active than ATP. Inhibition of enzyme activity by high levels of ATP was reversed by increasing the concentration of fructose-6-phosphate; the affinity of enzyme for fructose-6-phosphate decreased with increasing concentration of ATP. 5′-AMP, 3′,5′-AMP, 3′-AMP, deoxy AMP, UMP, IMP, CMP, GMP, ADP, CDP, GDP and UDP did not reverse the inhibition of enzyme by ATP. ADP, phosphoenolpyruvate and citrate inhibited phosphofructokinase activity but Pi did not affect it. Phosphofructokinase was not reactivated reversibly by mild change of pH and addition of effectors.  相似文献   

17.
Acetylcholine receptors were assayed with α-bugarotoxin on embryonic chick skeletal muscle growing in primary cell culture. Toxin was bound specifically to muscle cells and could be competed with D-tubocurarine. Two dissociation constants were obtained by equilibrium binding: 7.2 × 10?9M and 2.7 × 10?7M at 25°C. Two sets of rate constants were also obtained from dissociation kinetics. There are five times more low affinity sites on cells than high affinity sites. The average density of high-affinity receptors is about 200/μm2. A time course of toxin binding to receptors at 37°C vs 25°C in growth medium revealed that under conditions permitting growth and metabolism, toxin bound to cells was lost. The possibility that the growth medium was in-activating toxin molecules was ruled out by showing that unbound toxin molecules in the medium were fully capable of binding to fresh cultures.  相似文献   

18.
Abstract: Choline mustard aziridinium is a potent, irreversible and selective blocker of sodium-dependent, high-affinity transport of choline into rat forebrain synaptosomes; it was found to be 30 times less potent against low-affinity transport of choline. The IC50 value for high-affinity transport was 0.94 μM, compared to 29 μM for low-affinity uptake. The inhibitory action of choline mustard aziridinium ion on high-affinity transport of choline was graded with respect to time; a 12-fold increase in potency was obtained by increasing the inhibitor preincubation times from 1 to 30 min. Low concentrations of choline mustard aziridinium ion could produce significant blockade of choline carriers providing the exposure time was prolonged. The characteristics of the blockade of synaptosomal high-affinity choline transport by choline mustard aziridinium ion also changed depending upon preincubation time. The kinetics of inhibition of high-affinity choline transport by choline mustard aziridinium ion showed apparent competitive inhibition initially, followed by noncompetitive characteristics at longer preincubations with inhibitor. The rate of irreversible inhibition of carriers by this nitrogen mustard analogue would appear to be rapid; the rate constant was determined to be 5 × 10?2 s?1for micromolar concentrations of inhibitor. This action may preclude the transport of the mustard analogue into the nerve terminal, although initially some reversible binding with the carrier may result in the translocation of some choline mustard aziridinium ion into the presynaptic ending. The progressive alkylation of high-affinity carriers by the analogue could indicate the presence of excess carrier sites in the presynaptic membrane, or subpopulations of carriers in an inactive state in equilibrium with active carriers. A model is described for the inhibitory action of choline mustard aziridinium ion on synaptosomal high-affinity choline carriers.  相似文献   

19.
The kinetics of Cl-SO4-(2) exchange in Ehrlich ascites tumor cells was investigated in an attempt to determine the stoichiometry of this process. When tumor cells, equilibrated in Cl--free, 25 mM SO4-(2) medium are placed in SO4-(2)-free, 25 mm Cl-medium, both the net amount and rate of Cl-uptake far exceeds SO4-(2) loss.. Addition of the anion transport inhibitor SITS (4-acetamido-4,-isothiocyano-stilbene-2,2'-disulfonic acid) greatly reduces sulfate efflux (97%), but has no measurable effect on chloride uptake. Addition of furosemide, a Cl-transport inhibitor, reduces chloride uptake 94% but is without effect on sulfate efflux. These findings suggest that a chloride permeability pathway exists distinct from that utilized by SO4-(2). SITS, when added to furosemide treated cells, further reduces chloride uptake as well as inhibiting sulfate efflux, and under these experimental conditions, a linear relationship exists between SITS-sensitive, net chloride uptake and sulfate loss. The slope of this line is 1.05 (correlation coefficient = 0.996) which suggests the stoichiometry of Cl-SO4-(2) exchange is 1:1. Assuming a 1:1 stoichiometry, measurement of the initial chloride influx and initial sulfate efflux indicate that 92% of net chloride uptake is independent of sulfate efflux. Taken altogether, these results support the contention that the tumor cell possesses a permeability pathway which facilitates the exchange of one sulfate for one chloride. Under conditions where anion transport is not inhibited, this coupling is obscured by a second and quantitatively more important pathway for chloride uptake. This pathway is SITS-insensitive, although partially inhibited by furosemide.  相似文献   

20.
The modes of binding of 5′‐[4‐(aminoiminomethyl)phenyl]‐[2,2′‐Bifuran]‐5‐carboximidamide (DB832) to multi‐stranded DNAs: human telomere quadruplex, monomolecular R‐triplex, pyr/pur/pyr triplex consisting of 12 T*(T·A) triplets, and DNA double helical hairpin were studied. The optical adsorption of the ligand was used for monitoring the binding and for determination of the association constants and the numbers of binding sites. CD spectra of DB832 complexes with the oligonucleotides and the data on the energy transfer from DNA bases to the bound DB832 assisted in elucidating the binding modes. The affinity of DB832 to the studied multi‐stranded DNAs was found to be greater (Kass ≈ 107M?1) than to the duplex DNA (Kass ≈ 2 × 105M?1). A considerable stabilizing effect of DB832 binding on R‐triplex conformation was detected. The nature of the ligand tight binding differed for the studied multi‐stranded DNA depending on their specific conformational features: recombination‐type R‐triplex demonstrated the highest affinity for DB832 groove binding, while pyr/pur/pyr TTA triplex favored DB832 intercalation at the end stacking contacts and the human telomere quadruplex d[AG3(T2AG3)3] accommodated the ligand in a capping mode. Additionally, the pyr/pur/pyr TTA triplex and d[AG3(T2AG3)3] quadruplex bound DB832 into their grooves, though with a markedly lesser affinity. DB832 may be useful for discrimination of the multi‐sranded DNA conformations and for R‐triplex stabilization. © 2009 Wiley Periodicals, Inc. Biopolymers 93: 8–20, 2010. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  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