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 15 毫秒
1.
Cultured chondrocytes from the Swarm rat chondrosarcoma incorporate [35S]sulfate into proteoglycans typical of hyaline cartilage. The movement of newly synthesized proteoglycans from inside the cells into the extracellular matrix and, finally, into the culture medium was examined by measuring the distribution of 35S-labeled proteoglycans in the medium, a 4 M guanidine HCl extract of the cell layer, and in the remaining residue for a number of chase times following a 5-min pulse with [35S]sulfate. When hyaluronate oligosaccharides containing greater than or equal to 10 monosaccharides were included in the chase media, a proportion of newly synthesized proteoglycans were displaced from the matrix (4 M extract) into the culture medium. This displacement was greatest when oligomers were in the chase media between 10 and 20 min after the pulse, approximately the time when the molecules are being secreted from the cells. The proportion of link-stabilized aggregate in the medium was examined by Sepharose 2B chromatography after adding an excess of unlabeled monomer which displaces labeled monomer from complexes with hyaluronate which are not link-stabilized. The proportion of link-stabilized aggregate increased from 12% to about 70% between 12 and 120 min of chase. The presence of 40 micron hyaluronate oligosaccharides of 16 monosaccharides in the chase media retarded but did not prevent aggregate formation. Oligomers of about 50 monosaccharides, which are large enough to bind both a monomer proteoglycan and a link protein, almost completely prevented the formation of the large link-stabilized aggregates. The results suggest: (a) newly synthesized proteoglycans are not bound into link-stabilized aggregates at the time of secretion; (b) hyaluronic acid oligomers which are long enough to interact only with the hyaluronic acid-binding site of proteoglycans will retard but not prevent link-stabilized aggregation; and (c) hyaluronic acid oligomers long enough to accommodate additionally a link protein form a link-stabilized ternary complex and prevent aggregation with larger hyaluronic acid molecules.  相似文献   

2.
In aqueous solutions N-methyl-N-D-fructosyl amphotericin B methyl ester (MFAME), a novel amphotericin B derivative with low animal toxicity, similar to its parent antibiotic, exists in three forms: monomeric, soluble and insoluble aggregates in equilibrium [1]. The aim of our work was to examine the influence of medium composition on the MFAME self-association and the relationship between MFAME self-association and its toxicity towards red blood cells. The toxicity of MFAME in aggregated state towards red blood cells was tested by measuring the induction of potassium leakage and extent of haemolysis. The proportions of antibiotic species present in various aqueous media were determined by analysis of the UV-Vis spectra as a function of the antibiotic concentration. Numeric decomposition of the spectra allowed identification of four spectral species present in MFAME solutions: monomeric and three aggregated forms. Our results indicate that these aggregates, named type I, type II and type III, are different in terms of spectral properties, as well as effectiveness towards red blood cells. Soluble aggregate types I and III are the active forms of MFAME towards erythrocytes. The medium composition seems to be the main factor determining which type of antibiotic aggregate prevails in solution.  相似文献   

3.
The binding interaction between amphotericin B and human serum albumin (HSA) has been studied using surface plasmon resonance (SPR) spectroscopy combined with a fluorescence quenching method to confirm the binding kinetic results. In this paper, the SPR method used to study the drug-protein interaction has been described in detail. The association rate constant, dissociation rate constant and the equilibrium association constant of amphotericin B binding to HSA were obtained using this method. To confirm the feasibility of the SPR method, a fluorescence quenching method was performed to obtain the equilibrium constant. In order to obtain more accurate results, experiment design was used to optimize the fluorescence quenching process. The two equilibrium association constants obtained using the two methods were 4.017 x 10(4) M(-1) (SPR) and 3.656 x 10(4) M(-1) (fluorescence quenching method) respectively.  相似文献   

4.
S V Rudenko 《Biofizika》1986,31(1):59-63
It has been shown that structural rearrangements induced by glycerol in bilayer lipid membranes (BLM) containing cholesterol facilitate the transmembrane transport of amphotericin B molecules in the direction of glycerol gradient. The addition of amphotericin B to the same side with glycerol results in a change in bilayer selectivity from the cation to the anion one. Besides, the final conductivity is blocked by tetraethylammonium from the solution with no amphotericin B added. It testifies to the transport of amphotericin molecules to the opposite side of the membrane. The transport effect depends on the cholesterol content in bilayer, ionic strength of the medium and slightly depends on temperature. It is concluded that transport of amphotericin B in such conditions differs from the diffusive one and is due to the formation of intermediate lipid phases in the course of structural rearrangements of bilayers.  相似文献   

5.
The interaction between the polyene macrolide antibiotic, amphotericin B, and ergosterol in egg phosphatidylcholine multilayers was investigated using head group and acyl chain nitroxide spin-labelled phosphatidylcholine as probes. At physiological concentrations of less than 15 mol% sterol in egg phosphatidylcholine multilayers amphotericin B accumulates near the head group region until an amphotericin B : ergosterol ratio of approximately 0.7 is achieved. As the proportion of amphotericin B is increased above this value, formation of an acyl chain disordering complex occurs which has an approximate antibiotic:sterol ratio of unity. Dicetyl phosphate was used to increase the solubility of ergosterol past its normal limit in pure egg phosphatidylcholine (approximately 15 mol%). At concentrations of ergosterol higher than 15 mol% a complex of two ergosterol molecules and one amphotericin B was postulated when there was insufficient antibiotic to form a 1:1 complex.  相似文献   

6.
The complex formation of chlorhexidine digluconate (CHX-G(2)) with hydroxypropyl-β-cyclodextrin (HPβCD) was studied using NMR spectroscopy. The results revealed that this surfactant agent shows an monomer/aggregate equilibrium, which is dependent on the concentration of this drug. This equilibrium can be modified by the presence of HPβCD, which reduces the aggregation of the CHX-G(2) molecules. An inclusion process of the CHX-G(2) aromatic residue within the cyclodextrin cavity was confirmed by 2D ROESY spectroscopy. (1)H NMR titration studies of CHX-G(2) with HPβCD in D(2)O confirmed the formation of higher order complexes between CHX-G(2) and HPβCD. Moreover, the addition of HPβCD into CHX-G(2) solutions forms insoluble aggregates. Such insoluble aggregates may result in the stacking of CHX-G(2) molecules on the surface of the CHX-G(2):HPβCD complexes.  相似文献   

7.
The energetics of interaction and the type of aggregate structure in lateral assemblies of up to five gramicidin molecules in the beta 6.3 helical conformation at the air/water interface was calculated using conformational analysis procedures. It was found that within the aggregate two types of gramicidin interaction occur. One leading to a linear organization with a mean interaction energy between monomers of -6 kcal/mol and one in a perpendicular direction leading to a circularly organization with a lower mean interaction energy of -10 kcal/mol. Extrapolation towards larger gramicidin assemblies predicts that gramicidin itself could form tubular structures similar to those found in the gramicidin-induced HII phase. The tryptophans appear to play an essential role in the tubular organization of the gramicidin aggregate, since they determine the cone shape of the monomer and contribute to the structure of the monomer and oligomer by stacking interactions. These results, which are discussed in the light of experimental observations of gramicidin self-association in model membranes and the importance of the tryptophans for HII phase formation, further support the view (Killian, J.A. and De Kruijff, B. (1986) Chem. Phys. Lipids 40, 259-284) that gramicidin is a first example of a new class of hydrophobic polypeptides which can form cylindrical structures within the hydrophobic core of the membrane.  相似文献   

8.
Roflamycoin--a new channel-forming antibiotic   总被引:1,自引:0,他引:1  
Ion permeability of lipid bilayers was studied in the presence of a new antifungal pentaene antibiotic, roflamycoin, the structure of which differs considerably from that of the well-known polyene channel-former amphotericin B. Both of them, however, show the property of increasing the membrane permeability only in the case of sterol-containing membrane when added on both its sides. The conductance is strongly dependent on the concentration of the antibiotic in the solutions and of sterol in the membrane. Unlike the amphotericin B channels, roflamycoin channels are potential-dependent and have short lifetime (approx. 1 s) and high conductance (approx. 100 ps in 1 M KCl), which increases linearly with the salt concentration and is not blocked by the familiar blockers of amphotericin B channels. The two antibiotics seem to have a common mechanism of channel formation, viz. the formation starts from two semi-pores assembled in the opposite monolayers from several molecules of the antibiotic and sterol. However, the inner diameter of the roflamycoin channel is larger because of the different antibiotic-to-sterol ratio in the channel aggregate. It is believed that the difference in the ratio is due to the presence of the methyl group in the polyene chain of roflamycoin, and the considerable difference in lifetimes of the two types of channels depends on the terminal groups of the antibiotics.  相似文献   

9.
Ion permeability of lipid bilayers was studied in the presence of a new antifungal pentaene antibiotic, roflamycoin, the structure of which differs of which differs considerably from that of the well-known polyene channel-former amphotericin B. Both of them, however, show the property of increasing the membrane permeability only in the case of sterol-containing membrane when added on both its sides. The conductance is strongly dependent on the concentration of the antibiotic in the solutions and of sterol in the membrane. Unlike the amphotericin B channels, roflamycoin channels are potential-dependent and have short lifetime (approx. 1 s) and high conductance (approx. 100 ps in 1 M KCl), which increases linearly with the salt concentration and is not blocked by the familiar blockers of amphotericin B channels. The two antibiotics seem to have a common mechanism of channel formation, viz. the formation starts from two semi-pores assembled in the opposite monolayers from several molecules of the antibiotic and sterol. However, the inner diameter of the roflamycoin channel is larger because of the different antibiotic-to-sterol ratio in the channel aggregate. It is believed that the difference in the ratio is due to the presence of the methyl group in the polyene chain of roflamycoin, and the considerable difference in lifetimes of the two types of channels depends on the terminal groups of the antibiotics.  相似文献   

10.
11.
Interaction of filipin and amphotericin B with sterols in phosphatidylcholine membranes has been studied using various spin probes; epiandrosterone, cholestanone, phosphatidylcholine with 12-nitroxide or 5-nitroxide stearate attached to 2 position and also with tempocholine at the head group. Filipin caused increase in the fluidity of cholesterol-containing phosphatidylcholine membranes near the center, while it rather decreased the fluidity near the polar surface. On the other hand, amphotericin B did not apparently affect the fluidity. In the electron spin resonance spectrum of steroid spin probes in the antibiotic-containing membranes, both bound and free signals were observed and the association constant was calculated from the siganal intensity. In the binding of steroids with filipin, both 3 and 17 positions were involved, while the 17 position was less involved in the binding with amphotericin B. Phase change in the host membrane markedly affected the interaction of filipin with epiandrosterone probe. The bound fraction jumped from 0.4 to 0.8 on going to the crystalline state and increased further with decrease in temperature. The overall splitting of the bound signal also increased on lowering the temperature below phase transition. This change was attributed to aggregate formation of filipin-steroid complexes in the crystalline state. On the other hand, effect of phase transition was much smaller on the interaction of amphotericin B with the steroid probe.  相似文献   

12.
Several factors contributed to differences in intracellular composition between sugar-tolerant (osmophilic) and nontolerant species of yeast. One such factor was the difference in accumulation of those nonelectrolytes whose uptake was not dominated by vigorous metabolism. In such cases (lactose and glycerol), the sugar-tolerant species had a much lower capacity for the solute than did the nontolerant species. Sucrose uptake was consistently different between all sugar-tolerant strains on the one hand and all nontolerant strains on the other. The difference was attributable in part to metabolism of sucrose by the nontolerant yeasts. The major difference between the two types of yeast, however, was the presence of one or more polyhydric alcohols at high concentrations within each of the sugar-tolerant strains but none of the nontolerant strains. In most cases the major polyol was arabitol. The solute concentration (and, hence, water availability) of the growth medium affected both the amount of arabitol produced by Saccharomyces rouxii and the proportion retained by the yeast after brief washing with water at 0 C. When the yeast was suspended in a buffer at 30 C, the polyol leaked out at a slow, constant, reproducible rate. The polyene antibiotic amphotericin B caused rapid release of polyol by the yeast, the rate being proportional to amphotericin concentration. Contact of the yeast with glucose (1 mM) caused an extremely rapid ejection of polyol which lasted less than 40 s. Some implications of these results are discussed, as is the role of the polyol as a compatible solute in determining the water relations of the yeast.  相似文献   

13.
A new approach is proposed to investigate the metabolic perturbation induced by drugs in cells. The effects of various concentrations of amphotericin B on the aerobic [1-13C]glucose metabolism in glucose-grown repressed Saccharomyces cerevisiae cells were studied as a function of time using 13C-, 1H-NMR and biochemical methods. The 13C enrichment of different compounds such as ethanol, glycerol and trehalose were determined by 1H-NMR spectroscopy. In the absence of amphotericin B, glycerol diffuses slowly from the internal to the external medium, whereas in its presence this diffusion is greatly facilitated by the formation of pores in the cell membrane. Amphotericin B has been found to exert a marked influence on the glucose consumption and the production of all metabolites; for example, at 1 microM, the glucose consumption and the production of ethanol decrease while the production of glycerol and trehalose increases. The 13C relative enrichments of ethanol, glycerol and trehalose are almost the same with and without the drug. Thus it can be concluded that amphotericin B induces a large effect on the production of these compounds in the cytosol but shows no significant influence on the mechanism of their formation. Upon addition of glucose, all the amino acid concentrations decrease continuously with time; this effect is more pronounced in the presence of the drug. The ratio of the integrated resonances of glutamate (C2 + C3)/C4 reflects the activity of pyruvate carboxylase relative to citrate synthase rather than to pyruvate dehydrogenase. Without amphotericin B, this ratio (approximately 1.0) is practically constant upon addition of glucose which suggests that the activities of pyruvate carboxylase and citrate synthase are equivalent. By contrast, upon coaddition of 25 mM glucose and 1 microM amphotericin B, the glutamate C4 resonance remains virtually unchanged while that of glutamate C2 is much smaller than in its absence and continuously decreases with time. It seems likely that amphotericin B induces a reduction in the activity of pyruvate carboxylase in the mitochondria.  相似文献   

14.
In medium containing 40% dimethylsulfoxide, soluble F1 catalyzes the hydrolysis of ATP introduced at concentrations lower than that of the enzyme [Al-Shawi, M.K. & Senior, A.E. (1992), Biochemistry 31, 886-891]. At this concentration of dimethylsulfoxide, soluble F1 also catalyzes the spontaneous synthesis of a tightly bound ATP to a level of approximately 0.15 mol per mol F1 [Gómez-Puyou, A., Tuena de Gómez-Puyou, M. & de Meis, L. (1986) Eur. J. Biochem. 159, 133-140]. The mechanisms that allow soluble F1 to carry out these apparently opposing reactions were studied. The rate of hydrolysis of ATP bound to F1 under uni-site conditions and that of synthesis of ATP were markedly similar, indicating that the two ATP molecules lie in equivalent high affinity catalytic sites. The number of enzyme molecules that have ATP at the high affinity catalytic site under conditions of synthesis or uni-site hydrolysis is less than the total number of enzyme molecules. Therefore, it was hypothesized that when the enzyme was treated with dimethylsulfoxide, a fraction of the F1 population carried out synthesis and another hydrolysis. Indeed, measurements of the two reactions under identical conditions showed that different fractions of the F1 population carried out simultaneously synthesis and hydrolysis of ATP. The reactions continued until an equilibrium level between F1.ADP + Pi <--> F1.ATP was established. At equilibrium, about 15% of the enzyme population was in the form F1.ATP. The DeltaG degrees of the reaction with 0.54 microM F1, 2 mM Pi and 10 mM Mg2+ at pH 6.8 was -2.7 kcal.mol-1 in favor of F1.ATP. The DeltaG degrees of the reaction did not exhibit important variations with Pi concentration; thus, the reaction was in thermodynamic equilibrium. In contrast, DeltaG degrees became significantly less negative as the concentration of dimethylsulfoxide was decreased. In water, the reaction was far to the left. The equilibrium constant of the reaction diminished linearly with an increase in water activity. The effect of solvent is fully reversible. In comparison to other enzymes, F1 seems unique in that solvent controls the equilibrium that exists within an enzyme population. This results from the effect of solvent on the partition of Pi between the catalytic site and the medium, and the large energetic barrier that prevents release of ATP from the catalytic site. In the presence of dimethylsulfoxide and Pi, ATP is continuously hydrolyzed and synthesized with formation and uptake of Pi from the medium. This process is essentially an exchange reaction analogous to the phosphate-ATP exchange reaction that is catalyzed by the ATP synthase in coupled energy transducing membranes.  相似文献   

15.
The conditions of formation of amphotericin B-cholesterol or -ergosterol complexes in monolayers are investigated by the penetration into a monolayer of egg phosphatidylcholine/sterol of 14C-labelled N-fructosyl-amphotericin B dissolved in the aqueous subphase. An increase of both surface pressure and radioactivity as a function of concentration are observed simultaneously while a 'saturation' effect occurs only for the surface pressure. The experiments are not accurate enough to make conclusions about the number of actually penetrated amphotericin B molecules. Therefore, the existence of an amphotericin B-sterol complex was evidenced from a study of surface pressure area per molecule isotherm. The results indicate that a complex with a 2:1 stoichiometry is formed and that the amphotericin B-ergosterol interaction is larger than the amphotericin B-cholesterol interaction. The complex is dissociated by addition of egg phosphatidylcholine due to a competition between egg phosphatidylcholine and amphotericin B for sterol.  相似文献   

16.
The interaction of amphotericin B with ergosterol was studied in aqueous solutions of propanol. The mode of the interaction was found to be related to the aggregation state of amphotericin B. Ergosterol does not react (or reacts extremely slowly) with monomeric amphotericin B. Traces of a small aggregate, probably a dimer, enable a cooperative reaction. At high concentrations of the dimer, the reaction is immediate and the concentration of amphotericin B complexed with ergosterol is twice as high as the amount of added sterol. The interaction with ergosterol is hindered when the antibiotic is in micellar form. The pharmaceutical form, Fungizone, behaves similarly to the pure amphotericin B. Fungizone's greater solubility in water does not modify either the extent or the mode of interaction with ergosterol.  相似文献   

17.
The binding of polyene antibiotic amphotericin B to serum albumin was studied using absorption, fluorescence, and circular dichroism techniques. A hypochromic effect was observed in the absorption spectrum of amphotericin B in the presence of albumin with maxima at 366 nm, 385 nm, and 408 nm, which correspond to the absorption of the monomeric form of amphotericin B. A modification on the circular dichroism spectrum of amphotericin B in the presence of albumin was observed at bands 329 nm and 351 nm (excitronic interaction), which suggests that only amphotericin B monomer is bound to the protein. Amphotericin B perturbs the specific markers for sites I, II, and fatty acid binding site bound to these sites, suggesting that amphotericin B interacts with a great binding area in albumin. Lysines 199 and 525 in albumin participate in the molecular interaction between amphotericin B and the protein. The absorption spectrum of amphotericin B bound to albumin was sensitive to the chemical and thermal treatment of the protein, to neutral-basic transition of albumin and to conformational changes induced by the binding of other ligands to this protein.  相似文献   

18.
Amphiphilic vinyl polymers (in the free and cross-linked forms), carrying carboxyl and imidazole groups, were prepared by a radical polymerization of the purposely synthesized N-acryloyl-L-histidine. The protonation thermodynamic studies (at 25 degrees C in 0.15 M NaCl) showed high polyelectrolyte character of the soluble polymer. Unlike the linear decreasing trend of the basicity constant, over the whole range of alpha (degree of protonation), the enthalpy changes for the protonation of the imidazole nitrogen in the polymer showed a decreasing pattern only at alpha > 0.5. This was ascribed to the formation of hydrogen bonds between protonated and free neighboring monomer units. Viscometric data revealed a minimum hydrodynamic volume of the polymer at its isoelectric point (pH 5), whereas at higher or lower pHs, the macromolecule expanded greatly as a consequence of the charged sites formation. This produced a preferential solvation of the protonated imidazole and carboxylate ions, the latter being surrounded by more water molecules in the hydration shell. The peculiar hydration behavior was confirmed in the cross-linked polymer. The hydrogel showed an equilibrium degree of swelling (EDS), strongly dependent on pH, in a similar manner as viscometric data of the soluble polymer. A linear relationship between the reduced viscosity and the EDS was found. The polymer was non toxic against the RAW264 cell line.  相似文献   

19.
Change in aggregation state of insulin upon conjugation with 5-dimethylaminonaphthalene-1-sulfonyl (DNS) group was investigated at neutral pH. DNS group was introduced exclusively into B1 phenylalanine, the N-terminus of the B-chain of insulin. The association state of insulin shifted toward a more highly aggregated one upon conjugation, depending on the mole fraction (d) of DNS group to insulin monomer; at d equal 0.3 the equilibrium between dimer and hexamer was dominant over the range of 1-600 microM, while at d equal 1.0-1.5 DNS-insulin formed a larger aggregate (dodecamer) which is stable over the range of 67-600 microM. The dissociation constant of dimer-hexamer equilibrium at d=0.3 was evaluated to be 2.5 x 10(-10) M2 from the fluorescence anisotropy of the DNS group, which was about one order of magnitude smaller than that of the dimer-hexamer equilibrium in native insulin. Spectroscopic data and fluorescence decay analyses indicated that there exist at least two different environments surrounding the dye bound to B1 phenylalanine and that they are both relatively hydrophilic. It is considered that the major part of DNS group has excitation and emission maxima at longer wavelength with relatively low quantum yield, while the minor part has excitation and emission maxima at shorter wavelengths with relatively high quantum yield. The fluorescence lifetime of the dye was modified by the change in quaternary structure of DNS-lifetime of the dye was modified by the change in quaternary structure of DNS-insulin. Remarkable depolarization of DNS fluorescence was observed at d equal 1.0 and d equal 1.5 due to energy transfer between DNS groups conjugated to B1 phenylalanine in the hexamer or the dodecamer. Critical transfer distance for inter-DNS energy transfer was evaluated to be 15 A. From the molecular model of the insulin crystal, this energy transfer is ascribed to the close proximity, within about 15 A, between DNS groups in dimer units of the hexamer or the dodecamer.  相似文献   

20.
The recently discovered cyclotides kalata B1 and kalata B2 are miniproteins containing a head-to-tail cyclized backbone and a cystine knot motif, in which disulfide bonds and the connecting backbone segments form a ring that is penetrated by the third disulfide bond. This arrangement renders the cyclotides extremely stable against thermal and enzymatic decay, making them a possible template onto which functionalities can be grafted. We have compared the hydrodynamic properties of two prototypic cyclotides, kalata B1 and kalata B2, using analytical ultracentrifugation techniques. Direct evidence for oligomerization of kalata B2 was shown by sedimentation velocity experiments in which a method for determining size distribution of polydisperse molecules in solution was employed. The shape of the oligomers appears to be spherical. Both sedimentation velocity and equilibrium experiments indicate that in phosphate buffer kalata B1 exists mainly as a monomer, even at millimolar concentrations. In contrast, at 1.6 mm, kalata B2 exists as an equilibrium mixture of monomer (30%), tetramer (42%), octamer (25%), and possibly a small proportion of higher oligomers. The results from the sedimentation equilibrium experiments show that this self-association is concentration dependent and reversible. We link our findings to the three-dimensional structures of both cyclotides, and propose two putative interaction interfaces on opposite sides of the kalata B2 molecule, one involving a hydrophobic interaction with the Phe6, and the second involving a charge-charge interaction with the Asp25 residue. An understanding of the factors affecting solution aggregation is of vital importance for future pharmaceutical application of these molecules.  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