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.
Properties of mixed dispersions of sphingomyelin and the nonionic detergent, Triton X-305, were investigated by analytical ultracentrifugation and by autocorrelation spectroscopy of scattered laser light. These properties were compared with those of the sphingomyelin/Triton X-100 mixed micellar system reported previously [S. Yedgar, Y. Barenholz, and V. G. Cooper (1974) Biochim. Biophys. Acta 363, 98-111]. The substitution of the 30-unit ethylene oxide chain of Triton X-305 for the 10-unit chain of the Triton X-100 resulted in the appearance of two micellar phases at all detergent/lipid mixture ratios studied, whereas only a single mixed micellar phase was observed using Triton X-100. Despite this difference, the properties of the mixed lipid/detergent micelles obtained using Triton X-100 have been verified in the following respects: The detergent aggregation numbers in the mixed micelles are quite constant over a wide range of detergent molar fractions, being about 70 and 400 for the lighter and heavier mixed micellar phases, respectively. The detergent aggregation numbers are larger in the mixed micelle than in the pure detergent micelle. Very large sphingomyelin aggregation numbers can be accommodated within the mixed micelles, apparently by the critical intervention of the detergent molecules to produce a stable micellar structure.  相似文献   

2.
The mechanisms governing the solubilization by Triton X-100, octyl glucoside, and sodium cholate of large unilamellar liposomes prepared by reverse-phase evaporation were investigated. The solubilization process is described by the three-stage model previously proposed for these detergents [Lichtenberg, D., Robson, R.J., & Dennis, E.A.(1983) Biochim. Biophys. Acta 737, 285-304]. In stage I, detergent monomers are incorporated into the phospholipid bilayers until they saturate the liposomes. At that point, i.e., stage II, mixed phospholipid-detergent micelles begin to form. By stage III, the lamellar to micellar transition is complete and all the phospholipids are present as mixed micelles. The turbidity of liposome preparations was systematically measured as a function of the amount of detergent added for a wide range of phospholipid concentrations (from 0.25 to 20 mM phospholipid). The results allowed a quantitative determination of RSat, the effective detergent to lipid molar ratios in the saturated liposomes, which were 0.64, 1.3, and 0.30 for Triton X-100, octyl glucoside, and sodium cholate, respectively. The corresponding ratios in the mixed micelles, RSol, were 2.5, 3.8, and 0.9 mol of detergent/mol of phospholipid. The monomer concentrations of the three detergents in the aqueous phase were also determined at the lamellar to micellar transitions (0.18, 17, and 2.8 mM, respectively). These transitions were also investigated by 31P NMR spectroscopy, and complete agreement was found with turbidity measurements. Freeze-fracture electron microscopy and permeability studies in the sublytic range of detergent concentrations indicated that during stage I of solubilization detergent partitioning between the aqueous phase and the lipid bilayer greatly affects the basic permeability of the liposomes without significantly changing the morphology of the preparations. A rough approximation of the partition coefficients was derived from the turbidity and permeability data (K = 3.5, 0.09, and 0.11 mM-1 for Triton X-100, octyl glucoside, and sodium cholate, respectively). It is concluded that when performed systematically, turbidity measurements constitute a very convenient and powerful technique for the quantitative study of the liposome solubilization process by detergents.  相似文献   

3.
The human multidrug-resistance-associated protein 1 (hMRP1/ABCC1) belongs to the large ATP-binding cassette transporter superfamily. In normal tissues, hMRP1 is involved in tissue defense, whereas, in cancer cells, it is overproduced and contributes to resistance to chemotherapy. We previously investigated the folding properties of the predicted transmembrane fragments (TM) TM16, and TM17 from membrane-spanning domain 2 (MSD2). These TMs folded only partially as an α-helix and were located in the polar headgroup region of detergent micelles used as membrane mimics (Vincent et al. in Biochim Biophys Acta 1768:538–552, 2007; de Foresta et al. in Biochim Biophys Acta 1798:401–414, 2010). We have now extended these studies to TM4 and TM10, from MSD0 and MSD1, respectively. TM10 may be involved in the substrate translocation pathway whereas the role of TM4 is less predictable, because few studies have focused on MSD0, a domain present in some hMRP1 homologs only. Each TM contained a single Trp residue (W142 or W553) acting as an intrinsic fluorescent probe. The location and dynamics of the TMs in dodecylphosphocholine (DPC) or n-dodecyl-β-d-maltoside (DDM) micelles were studied by Trp steady-state and time-resolved fluorescence, including quenching experiments. Overall TM structure was analyzed by far-UV circular dichroism studies in detergent micelles and TFE. TM10 behaved similarly to TM16 and TM17, with an interfacial location in micelles consistent with a possible role in lining the transport pore. By contrast, TM4 behaved like a classical TM fragment with a high α-helical content, and its transmembrane insertion did not require its interaction with other TMs.  相似文献   

4.
In this paper we present the results of a Monte Carlo study of the effects of protein, cholesterol, bilayer curvature, and mobility on the chain order parameters of a lipid layer. The Monte Carlo method used is identical to the version developed earlier (Scott, Jr., H.L. (1977) Biochim. Biophys. Acta 469, 264–271). Simulations of protein and cholesterol effects are accomplished by insertion of a rigid stationary cylinder into the lipid matrix. The protein studies show the presence of boundary lipid (Jost, P., Griffith, O.H., Capaldi, R.H. and Vanderkooi, G. (1973) Biochim. Biophys. Acta 311, 141–152). The effect of cholesterol is dependent upon the length of the lipid hydrocarbon chains relative to the cholesterol depth of penetration. Our computer studies of bilayer curvature show the manner in which this curvature disrupts chain packing and are consistent with experimental results (Chrzeszczyk, A., Wishnia, A. and Springer, C.S. (1977) Biochim. Biophys. Acta, 470, 161–171). We also find that restricting lateral motion in chains, the simplest manner in which head group interactions can affect hydrocarbon chain order, does not measurably alter the order parameters. We argue that this provides some support for an earlier hypothesis by Scott (Scott, Jr., H.L. (1975) Biochim. Biophys. Acta 406, 329–346) regarding head group-chain interaction in monolayer experiments.  相似文献   

5.
Rat epididymal fat cell membrane proteins were extracted from adipocyte ghosts with octylglucoside and incorporated by detergent dialysis into unilamellar phosphatidylcholine vesicles approx. 200 nm in diameter. The rate of glucose transport into the vesicles under zero-trans conditions was substrate dependent, saturable and inhibited by phloretin and cytochalasin B. Their maximum specific transport activity was 35.6 mumol/min per mg protein, and their half saturation constant for glucose was 15 mM. Glucose transport into the reconstituted vesicles was inhibited by only those sugars which competitively inhibited glucose transport into intact adipocytes. A major protein component of the vesicles was a 100 kDa protein which we had previously found to react with the affinity label maltosyl isothiocyanate (Malchoff, D.M., Olansky, L., Pohl, S. and Langdon, R.G. (1981) Fed. Proc. 40, 1893). Removal of adipocyte ghost membrane extrinsic proteins with dimethylmaleic anhydride followed by extraction of the resulting membrane pellet with octylglucoside yielded a solution which contained two major proteins, of Mr 100 000 and 85 000, with very small quantities of lower Mr proteins. Vesicles into which these proteins were incorporated had average specific transport activities of 624 mumol/min per mg protein and half saturation constants of 22 mM. Our results strongly indicate that the native glucose transporter of the rat adipocyte, like that of the human erythrocyte (Shelton, R.L. and Langdon, R.G. (1983) Biochim. Biophys. Acta 733, 25-33), is a 100 kDa protein.  相似文献   

6.
A mutant of Escherichia coli missing energy-transducing ATPase and known to be defective in a variety of membrane functions from earlier studies (Yamamoto, T. H., Mével-Ninio, M. and Valentine, R. C. (1973) Biochim. Biophys. Acta 314, 267-275; Thipayathasana, P. and Valentine, R. C. (1974) Biochim. Biophys. Acta 347, 464-468; Mével-Ninio, M. and Yamamoto, T. (1974) Biochim. Biophys. Acta 357, 63-66) has been found to be blocked for anaerobic DNA synthesis. The rate of anaerobic DNA synthesis in the mutant, measured as radioactive adenine incorporation into the alkali-resistant fraction of whole cells, is about 1/6 the rate of DNA synthesis in the wild type culture under similar conditions. Addition of NO-3- or O-2 restores DNA biosynthesis in the mutant. The entry of radioactive adenine is not appreciably affected in the mutant by anaerobiosis. It is concluded that coupling factor plays a role in some step(s) of DNA biosynthesis.  相似文献   

7.
The kinetic features of glucose transport in human erythrocytes have been the subject of many studies, but no model is consistent with both the kinetic observations and the characteristics of the purified transporter. In order to reevaluate some of the kinetic features, initial rate measurements were performed at 0 degree C. The following kinetic parameters were obtained for fresh blood: zero-trans efflux Km = 3.4 mM, Vmax = 5.5 mM/min; infinite-trans efflux Km = 8.7 mM, Vmax = 28 mM/min. For outdated blood, somewhat different parameters were obtained: zero-trans efflux Km = 2.7 mM, Vmax = 2.4 mM/min; infinite-trans efflux Km = 19 mM, Vmax = 23 mM/min. The Km values for fresh blood differ from the previously reported values of 16 mM and 3.4 mM for zero-trans and infinite-trans efflux, respectively (Baker, G.F. and Naftalin, R.J. (1979) Biochim. Biophys. Acta 550, 474-484). The use of 50 mM galactose rather than 100 mM glucose as the infinite-trans sugar produced no change in the infinite-trans efflux Km values but somewhat lower Vmax values. Simulations indicate that initial rates were closely approximated by the experimental conditions. The observed time courses of efflux are inconsistent with a model involving rate-limiting dissociation of glucose from hemoglobin (Naftalin, R.J., Smith, P.M. and Roselaar, S.E. (1985) Biochim. Biophys. Acta 820, 235-249). The results presented here support the adequacy of the carrier model to account for the kinetics.  相似文献   

8.
β2-GlycoproteinI was shown to be a hydrophilic protein exhibiting no charge shift in the presence of a cationic detergent, but a charge shift in the presence of an anionic detergent. The latter was suggested to be caused by a binding of β2-glycoprotein I to deoxycholate in the Triton X-100/deoxycholate micelles. Quantification by electroimmunoassay of asialo-β2-glycoprotein and I and subfractions of β2-glycoprotein I gave different results although identical results were obtained in single radial immunodiffusion. Addition of 0.2% (w/v) of deoxycholate to the agarose gels containing Triton X-100 prior to electrophoresis, however, eliminated these differences. The effect of deoxycholate on the rate of migration of β2-glycoprotein I was found applicable for electroimmunoassay of the protein. Crossed immunofocusing of plasma from individual donors, electrophoresed in an agarose containing Triton X-100/deoxycholate micelles confirmed a postulated variation in the relative composition of subfractions of β2-glycoprotein I in plasma earlier suggested by Finlayson and Mushinski (Finlayson, J.S. and Mushinski, J.F. (1967) Biochim. Biophys. Acta 147, 413–420).  相似文献   

9.
A mutant of Escherichia coli missing energy-transducing ATPase and known to be defective in a variety of membrane functions from earlier studies (Yamamoto, T. H., Mével-Ninio, M. and Valentine, R. C. (1973) Biochim. Biophys. Acta 314, 267–275; Thipayathasana, P. and Valentine, R. C. (1974) Biochim. Biophys. Acta 347, 464–468; Mével-Ninio, M. and Yamamoto, T. (1974) Biochim. Biophys. Acta 357, 63–66) has been found to be blocked for anaerobic DNA synthesis. The rate of anaerobic DNA synthesis in the mutant, measured as radioactive adenine incorporation into the alkali-resistant fraction of whole cells, is about 1/6 the rate of DNA synthesis in the wild type culture under similar conditions. Addition of NO3- or O2 restores DNA biosynthesis in the mutant. The entry of radioactive adenine is not appreciably affected in the mutant by anaerobiosis. It is concluded that coupling factor plays a role in some step(s) of DNA biosynthesis.  相似文献   

10.
We examined the action of porcine pancreatic and bee-venom phospholipase A2 towards bilayers of phosphatidylcholine as a function of several physical characteristics of the lipid-water interface. 1. Unsonicated liposomes of dimyristoyl phosphatidylcholine are degraded by both phospholipases in the temperature region of the phase transition only (cf. Op den Kamp et al. (1974) Biochim. Biophys. Acta 345, 253--256 and Op den Kamp et al. (1975) Biochim. Biophys. Acta 406, 169--177). With sonicates the temperature range in which hydrolysis occurs is much wider. This discrepancy between liposomes and sonicates cannot be ascribed entirely to differences in available substrate surface. 2. Below the phase-transition temperature the phospholipases degrade dimyristoyl phosphatidylcholine single-bilayer vesicles with a strongly curved surface much more effectively than larger single-bilayer vesicles with a relatively low degree of curvature. 3. Vesicles composed of egg phosphatidylcholine can be degraded by pancreatic phospholipase A2 at 37 degrees C, provided that the substrate bilayer is strongly curved. The bee-venom enzyme shows a similar, but less pronounced, preference for small substrate vesicles. 4. In a limited temperature region just above the transition temperature of the substrate the action of both phospholipases initially proceeds with a gradually increasing velocity. This stimulation is presumably due to an increase of the transition temperature, effectuated by the products of the phospholipase action. 5. Structural defects in the substrate bilayer, introduced by sonication below the phase-transition temperature (cf. Lawaczeck et al. (1976) Biochim. Biophys. Acta 443, 313--330) facilitate the action of both phospholipases. The results lead to the general conclusion that structural irregularities in the packing of the substrate molecules facilitate the action of phospholipases A2 on phosphatidylcholine bilayers. Within the phase transition and with bilayers containing structural defects these irregularities represent boundaries between separate lipid domains. The stimulatory effect of strong bilayer curvature can be ascribed to an overall perturbation of the lipid packing as well as to a change in the phase-transition temperature.  相似文献   

11.
The ADP/ATP carrier (AAC) from yeast mitochondria has been reconstituted in phospholipid vesicles essentially according to the procedure described for the reconstitution of AAC from bovine heart mitochondria (Kr?mer and Heberger (1986) Biochim. Biophys. Acta, 863, 289-296). Liposomes were prepared from the mixed micelles of dodecyl octaoxyethylene ether (C12E8)-solubilized protein and egg yolk phosphatidylcholine by removing the detergent with Amberlite treatment. The micelles were treated with Amberlite either by repeatedly passing through small columns filled with Amberlite XAD-2 beads or by stepwise addition of Amberlite beads to the micelles. All the important variables of the reconstitution components were kept at optimal level and the liposomes obtained by both the methods of Amberlite treatment were analysed for (3H)CAT binding, orientation of AAC and nucleotide exchange activity. Reconstituted AAC showed 80% right side out orientation in the liposomes prepared by either procedure. Lipsomes prepared by the Amberlite column procedure exhibited higher CAT binding but lower ADP exchange activity. Liposome preparation by the stepwise addition of Amberlite is suggested to be the method of choice for functional reconstitution of yeast AAC in view of the higher nucleotide transport activity associated with the liposomes prepared by this method.  相似文献   

12.
A water-soluble Mg2+-ATPase previously reported (White, M.D. and Ralston, G.B. (1976) Biochim. Biophys. Acta 436, 567-576) has been purified from human erythrocyte membranes. The purified enzyme has a molecular weight of 575 000; the apparent minimum molecular weight was 100 000, corresponding to a soluble protein of the component 3 region. The Km value for ATP was 1 mM and apparent Km for Mg2+ was 3.6 mM. By means of histochemical activity staining in acrylamide gels it was shown that the purified ATPase preparation could be inhibited by Cd2+ and Zn2+ salts, p-chloromercuribenzoate and N-ethylmaleimide, known inhibitors of membrane endocytosis.  相似文献   

13.
Reaction centers of the photosynthetic bacterium Rhodopseudomonas sphaeroides R-26, give rise to large triplet state EPR signals upon illumination at low temperature (11 K). Utilizing monochromatic polarized light to generate the EPR spectra (magnetophotoselection) we have shown that the intensities of the observed triplet signals are strongly dependent upon the wavelength and polarization direction of the excitation. These data can be used to calculate the orientations of the excited transition moments with respect to each other and with respect to the triplet state principal magnetic axes system. Our quantitative approach is to follow the procedure outlined in a previous publication (Frank, H.A., Friesner, R., Nairn, J.A., Dismukes, G.C. and Sauer, K. (1979) Biochim. Biophys. Acta 547, 484-501) where computer simulations of the observed triplet state spectra were employed. The results presented in the present work indicate that the transition moment at 870 nm which is associated with the bacteriochlorophyll 'special pair' lies almost entirely along one of the principal magnetic axes of the triplet state. Aso, the 870 nm transition moment makes an angle of approx. 60 degrees with the 546 nm transition moment which is associated with a bacteriopheophytin. This latter result is in agreement with previous photoselection studies on the same bacterial species (Vermeglio, A., Breton, J., Paillotin, G. and Cogdell, R. (1978) Biochim. Biophys. Acta 501, 514-530).  相似文献   

14.
The effect of a series of n-alcohols on the permeability of small, unilamellar dipalmitoyl phosphatidylcholine (DPPC), dimyristoyl phosphatidylcholine (DMPC) and distearoyl phosphatidylcholine (DSPC) vesicles at the gel-to-liquid crystal phase transition temperature was investigated. It was found that the permeability took the form of the transient lysis of a fraction of the population of vesicles. The effect on this lysis of the n-alcohols was seen to be very chain-length dependent, with a minimum at n = 8 (octan-1-ol) for DPPC vesicles. A similar minimum was observed in the presence of 0.1 mM Triton X-100, but the detergent could then interact with certain of the alcohols to produce permanent channels. The results are discussed in terms of the semi-empirical model of Brasseur et al. (1985) Biochim. Biophys. Acta 814, 227-236, for the interaction of the n-alcohols with a DPPC membrane. The effect of various n-alcohols on the outer and inner monolayers of DPPC vesicles was also studied and the results related to their fluidising effect, allowing channels to open at the phase transition temperature.  相似文献   

15.
X-irradiation of the ternary complex of P450:substrate:O2 at 77 K produces a reduced intermediate by electron addition to the Fe:O2 complex which can be studied by EPR-spectroscopy. The EPR spectrum of the new species exhibits rhombic symmetry with g-factors of 2.27, 2.17 and 1.95, respectively. Increasing the temperature of the sample to 190 K results in loss of intensity of the intermediate signals. X-irradiation of oxymyo- and oxyhemoglobin produces similar EPR signals indicating that the added electron is resident on the Fe:O2 compleX (Kappl, R., et al. (1985) Biochim. Biophys. Acta 870, 20-30).  相似文献   

16.
In 1972, Lieb, W. R. and Stein, W. D. (Biochim, Biophys. Acta 265, 187-207) in their review of sugar transport in human erythrocytes concluded that the conventional two-state carrier model was inconsistent with the experimental data available at that time. Since then, other papers have appeared which question the validity of the model. In this paper, we give a brief derivation of the equations describing the two-state carrier model, and analyze the predictions of the model in the classical experiments, i.e. zero-trans, infinite-cis, and equilibrium exchange. We show that the estimate of the half saturation constant of 2.8 mM for glucose at the inner face of the human red cell membrane for the infinite-cis procedure reported by Hankin, B.L., Liev, W.R. and Stein, W.D ((1972) Biochim. Biophys. Acta 288, 114-126) is unreliable. We note that all of the other experimental findings are consistent with the asymmetric carrier model.  相似文献   

17.
L M Peerey  N M Kosti? 《Biochemistry》1989,28(4):1861-1868
Horse heart cytochrome c and French bean plastocyanin are cross-linked one-to-one by a carbodiimide [Geren, L. M., Stonehuerner, J., Davis, D. J., & Millett, F. (1983) Biochim. Biophys. Acta 724, 62] in the same general orientation in which they associate electrostatically [King, G. C., Binstead, R. A., & Wright, P. E. (1985) Biochim. Biophys. Acta 806, 262]. The reduction potentials of the Fe and Cu atoms in the covalent diprotein complex are respectively 245 and 385 mV vs NHE; the EPR spectra of the two metals are not perturbed by cross-linking. Four isomers of the covalent diprotein complex, which probably differ slightly from one another in the manner of cross-linking, are separated efficiently by cation-exchange chromatography. Stopped-flow spectrophotometric experiments with the covalent diprotein complex show that the presence of plastocyanin somewhat inhibits oxidation of ferrocytochrome c by [Fe(CN)6]3- and somewhat promotes oxidation of this protein by [Fe(C5H5)2]+. These changes in reactivity are explained in terms of electrostatic and steric effects. Pulse-radiolysis experiments with the electrostatic diprotein complex yield association constants of greater than or equal to 5 X 10(6) and 1 X 10(5) M-1 at ionic strengths of 1 and 40 mM, respectively, and the rate constant of 1.05 X 10(3) s-1, regardless of the ionic strength, for the intracomplex electron-transfer reaction. Analogous pulse-radiolysis experiments with each of the four isomers of the covalent diprotein complex, at ionic strengths of both 2 and 200 mM, show an absence of the intracomplex electron-transfer reaction. A rearrangement of the proteins for this reaction seems to be possible (or unnecessary) in the electrostatic complex but impossible in the covalent complex.  相似文献   

18.
Guanylate cyclase activity was determined in a 1000g particulate fraction derived from rabbit heart homogenates using Mg2+ or Mn2+ as sole cation in the presence and absence of Triton X-100. With Mg2+, very little guanylate cyclase activity could be detected in the original particulate fraction assayed with or without Triton, or in the particulate fraction treated with varying concentrations of Triton (detergent-treated mixture) prior to enzyme assay. However, the detergent-solubilized supernatants as well as the detergent-insoluble residues (pellets) derived from detergent-treated mixtures possessed appreciable Mg2+-supported enzyme activity. With Mn2+, significant enzyme activity was detectable in the original particulate fraction assayed without Triton. Much higher activity was seen in particulate fraction assayed with Triton and in detergent-treated mixtures; the supernatants but not the pellets derived from detergent-treated mixtures possessed even greater activity. The sum of enzyme activity in pellet and supernatant fractions greatly exceeded that of the mixture. When the pellets and supernatants derived from detergenttreated mixtures were recombined, measured enzyme activities were similar to those of the original mixture. With Mg2+ or Mn2+, the specific activity of guanylate cyclase in pellet and supernatant fractions varied considerably depending on the concentration of Triton used for treatment of the particulate fraction; treatment with low concentrations of Triton (0.2–0.7 μmol/mg protein) gave supernatants showing high activity whereas treatment with relatively greater concentrations of the detergent (>0.7 μmol/mg protein) gave pellets showing high activity. The relative distribution of guanylate cyclase in pellet and supernatant fractions expressed as a function of Triton concentration during treatment (of the particulate fraction) showed that 50 to 80% of the recovered enzyme activity remained in supernatants at low detergent concentrations whereas 50 to 80% of the recovered activity resided in the pellets at higher detergent concentrations. Inclusion of excess Triton in the enzyme assay medium did not alter the specific activity profiles and the relative distribution patterns of the cyclase in pellet versus supernatant fractions. The results demonstrate the inherent potential of cardiac particulate guanylate cyclase to utilize Mg2+ in catalyzing the synthesis of cyclic GMP. However, it appears that some factor(s) endogenous to the cardiac particulate fraction severely impairs the expression of Mg2+-dependent activity; Mn2+-dependent activity is also affected by such factor(s) but apparently less severely. Further, the results suggest that previously reported activities of cardiac particulate guanylate cyclase, despite being assayed with Mn2+ and in the presence of Triton X-100, represent underestimation of what otherwise appears to be a highly active enzyme system capable of utilizing physiologically relevant divalent cation such as Mg2+.  相似文献   

19.
Reaction centers of the photosynthetic bacterium Rhodopseudomonas sphaeroides R-26, give rise to large triplet state EPR signals upon illumination at low temperature (11 K). Utilizing monochromatic polarized light to generate the EPR spectra (magnetophotoselection) we have shown that the intensities of the observed triplet signals are strongly dependent upon the wavelength and polarization direction of the excitation. These data can be used to calculate the orientations of the excited transition moments with respect to each other and with respect to the triplet state principal magnetic axes system. Our quantitative approach is to follow the procedure outlined in a previous publication (Frank, H.A., Friesner, R., Nairn, J.A., Dismukes, G.C. and Sauer, K. (1979) Biochim. Biophys. Acta 547, 484–501) where computer simulations of the observed triplet state spectra were employed.The results presented in the present work indicate that the transition moment at 870 nm which is associated with the bacteriochlorophyll ‘special pair’ lies almost entirely along one of the principal magnetic axes of the triplet state. Also, the 870 nm transition moment makes an angle of approx. 60° with the 546 nm transition moment which is associated with a bacteriopheophytin. This latter result is in agreement with previous photoselection studies on the same bacterial species (Vermeglio, A., Breton, J., Paillotin, G. and Cogdell, R. (1978) Biochim. Biophys. Acta 501, 514–530).  相似文献   

20.
A protein activator of glucosylceramidase (EC 3.2.1.45) has been previously identified by us in human placenta [(1985) Biochim. Biophys. Acta 836, 157-166]. In the present paper we report that its function in vitro is to stimulate the binding of the enzyme to its substrate, glucosylceramide. After the purification step which frees the enzyme of most of its activator protein (octyl-Sepharose 4B chromatography), the capacity of glucosylceramidase to bind to the glucosylceramide micelles is dramatically decreased. The addition of the activator protein to the purified enzyme restores this binding.  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