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.
Various models for the analysis of time-dependent fluorescence anisotropy measurements were evaluated. The discussion was based on the analysis of pulsed experiments with 1,6-diphenyl-1,3,5-hexatriene embedded in small unilamellar vesicles of dimyristoylphosphatidylcholine or dipalmitoylphosphatidylcholine and in dimyristoylphosphatidylcholine/alpha-lactalbumin complexes. It was shown that a recently proposed model (Van der Meer, W., H. Pottel, W. Herreman, M. Ameloot, H. Hendrickx, H. Schröder, 1984, Biophys. J., 46:515-523) described the data better than did the earlier suggested cone model (Kinosita K., Jr., S. Kawato, and A. Ikegami, 1977, Biophys. J., 20:289-305). This permitted the use of the new model for the estimation of the second- and fourth-rank order parameters on nonoriented systems. The results indicated that a fraction of the probes was oriented perpendicularly to the preferred direction of the lipids. An increase of the rotational correlation times of the fluorescent probe and a higher order of its environment were detected after the interaction of alpha-lactalbumin with the dimyristoylphosphatidylcholine vesicles at acidic pH at 24.2 degrees C.  相似文献   

2.
The dynamics of lipid hydrocarbon chains in phosphatidylcholine (dimyristoyl- or dipalmitoyl-) and cholesterol/dimyristoylphosphatidylcholine membranes were investigated by nanosecond time-resolved fluorescence depolarization measurements on a lipophilic fluorescent probe 1,6-diphenyl-1,3,5-hexatriene embedded in the membranes. In the pure lipid membranes, both the range (amplitude) and the rate of the wobbling motion of the probe increased sigmoidally with temperature reflecting the thermotropic phase transition of the lipid. The rise in the rate slightly preceded the increase in the range, suggesting that the fluctuation of lipid chains is activated to a high level before the ordered array of chains melt into the liquid-crystalline phase. Above the transition temperature, incorporation of cholesterol resulted in a dramatic decrease in the range of wobbling motion while the rate remained high. Below the transition, on the other hand, cholesterol had little effect on the range, whereas the rate was greatly increased. These effects of cholesterol are remarkably similar to the effects of cytochrome oxidase on lipid chain dynamics (Kinosita, K., Jr., Kawato, S., Ikegami, A., Yoshida, S. and Orii, Y. (1981) Biochim. Biophys. Acta 647, 7–17).  相似文献   

3.
Internal motion of the DNA in chicken erythrocyte chromatin fibers was studied by measurement of the fluorescence anisotropy decay of ethidium intercalated in the linker region. A comparison of the decay curves of the dye in chicken erythrocyte chromatin with those of calf thymus chromatin [Ashikawa, I., Kinosita, K., Jr., Ikegami, A., Nishimura, Y., Tsuboi, M., Watanabe, K., Iso, K., & Nakano, T. (1983) Biochemistry 22, 6018-6026] revealed greater suppression of nucleosome movement in chicken erythrocyte chromatin. Furthermore, the transition of this chromatin to the compact (solenoidal) structure occurred at lower solvent concentrations of Na+ or Mg2+ than those for calf thymus chromatin. These results demonstrated increased stability of the higher order structure (the solenoid) of chicken erythrocyte chromatin, which may be related to the reduction of nuclear activity in the chicken erythrocyte cell. In addition to intact chicken erythrocyte chromatin, we studied the structural transitions of H1-depleted and H1,H5-depleted chromatins. The result indicated that histone H5 of this chromatin stabilizes the higher order structure in the presence of magnesium (or divalent) cation and did not induce the transition in the solution containing only sodium cation.  相似文献   

4.
L Powers  B M Kincaid 《Biochemistry》1989,28(10):4461-4468
Differences in the methods of analysis of X-ray absorption data used by Powers et al. [Powers, L., Blumberg, W. E., Chance, B., Barlow, C., Leigh, J., Jr., Smith, J., Yonetani, T., Vik, S., & Peisach, J. (1979) Biochim. Biophys. Acta 547, 520-538; Powers, L., Chance, B., Ching, Y., & Angiolillo, P. (1981) Biophys. J. 34, 465-498] and Scott et al. [Scott, R., Schwartz, J., & Cramer S. (1986) Biochemistry 25, 5546-5555] are clarified. In addition, we compare the X-ray absorption data and results for resting cytochrome c oxidase reported by both groups using the same analysis method and conclude apart from any assumptions that the data are not identical.  相似文献   

5.
J Katsaras  D S Yang    R M Epand 《Biophysical journal》1992,63(4):1170-1175
X-ray diffraction has been applied to determine the various tilt angles and directions (if any) which can be assumed by oriented gel phase multilayers of dipalmitoyl phosphatidylcholine (DPPC) as a function of hydration. We report for the first time that oriented DPPC multilayers with a repeat spacing (d-spacing) of 55.2A at 25 degrees C and 0% relative humidity (RH) have hydrocarbon chains tilted at an angle theta of 21.5 degrees with respect to the bilayer normal. In addition, the chains are tilted along one of the bisectors (omega = 0 degrees) of the hexagonal lattice (8 wide-angle maxima, 2 unique), a phase not previously reported in DPPC studies. At 100% RH, the chain tilt angle and d-spacing increased to approximately 29.0 degrees and 58.9A, respectively. Since at 100% RH only 4 wide-angle maxima are observed, we analyze the data on the assumption that the hydrocarbon chains may rotate independently of the hexagonal lattice (omega = 0-30 degrees), at a fixed chain tilt angle theta (Stamatoff, J.B., et al. 1979. Biophys. J. 25:253-262). The largest observed angle phi made by the wide-angle maxima with the equator is 29.5 degrees corresponding to a theta of approximately 32.6 degrees (omega avg. = 24 degrees) and the sample having a d-spacing of 64.0 A (excess water condition). Finally, theta remains relatively constant (approximately 21.5 degrees) up to a RH of approximately 45% and a d-spacing of 57.8A, after which, with increases in RH, theta increases to a maximum of 32.6 degrees.  相似文献   

6.
The presence of pyrroloquinoline quinone (PQQ) as the organic cofactor of Dactylium dendroides galactose oxidase and lentil (Lens culinaris) seedling amine oxidase, purported PQQ-containing oxidoreductases (Van der Meer, R. A., Jongejan, J. A., and Duine, J. A. (1989) J. Biol. Chem. 264, 7792-7794; Citro, G., Verdina, A., Galati, R., Floris, G., Sabatini, S., and Finazzi-Argo', A. (1989) FEBS Lett. 247, 201-204), was reinvestigated using the nitro blue tetrazolium redoxcycling method (Paz, M. A., Gallop, P. M., Torrelio, B. M., and Flückiger, R. (1988) Biochem. Biophys. Res. Commun. 154, 1330-1337; Paz, M. A., Flückiger, R., Boak, A., Kagan, H. M., and Gallop, P. M. (1991) J. Biol. Chem. 266, 689-692) and the enzyme-linked immunosorbent assay with polyclonal antibodies against PQQ. The possible quinoprotein nature of the laccases from Polyporus versicolor and Rhus vernicifera was also investigated because of the similarities in spectroscopic and kinetic features of these enzymes and the laccase from Phlebia radiata, reported to be a PQQ protein (Karhunen, E., Niku-Paavola, M.-L., Viikari, L., Haltia, T., Van der Meer, R. A., and Duine, J. A. (1990) FEBS Lett. 267, 6-8). The presence of a quinonoid cofactor in lentil seedling amine oxidase is confirmed, whereas galactose oxidase and both laccases do not display any quinoprotein nature.  相似文献   

7.
Considering the effects of pH on many aspects of cell metabolism, such as its role in signaling processes and enzyme kinetics, it is indispensable to include the measurement of the dynamics of the intracellular pH, when studying the fast dynamic response of cells to perturbations. It has been shown previously that the intracellular pH rapidly drops following an increase in external glucose concentration [Kresnowati, M.T.A.P., Suarez-Mendez, C., Groothuizen, M.K., Van Winden, W.A., Heijnen, J.J., 2007. Measurement of fast dynamic intracellular pH in Saccharomyces cerevisiae using benzoic acid pulse. Biotechnol. Bioeng. 97, 86-98; Ramos, S., Balbin, M., Raposo, M., Valle, E., Pardo, L.A., 1989. The mechanism of intracellular acidification induced by glucose in Saccharomyces cerevisiae. J. Gen. Microbiol. 135, 2413-2422; Van Urk, H., Schipper, D., Breedveld, G.J., Mak, P.R., Scheffers, W.A., Van Dijken, J.P., 1989. Localization and kinetics of pyruvate-metabolizing enzymes in relation to aerobic alcoholic fermentation in Saccharomyces cerevisiae CBS 8066 and Candida utilis CBS 621. Biochim. Biophys. Acta 992(1), 78-86]. The mechanism for this fast intracellular acidification, however, has not been elucidated yet. This paper presents a metabolome-based analysis to reveal the physiological phenomena that cause the fast intracellular acidification following either a glucose pulse or an ethanol pulse to carbon-limited chemostat cultures of Saccharomyces cerevisiae. This quantitative study, which includes the determination of intracellular buffering capacity, the calculation of electric charge balance and the quantification of weak organic acid transport shows that none of the previously suggested mechanisms, i.e. increase in glucose phosphorylation and accumulation of CO(2), is sufficient to explain the measured decrease in intracellular pH following a glucose pulse.  相似文献   

8.
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   

9.
The amino-terminal sequence (33 residues) of the acid protease from Penicillium roqueforti has been determined with an automated sequencer. The amino-terminal sequence of Rhizopus pepsin (published by Sepulveda, P., Jackson, K. W. & Tang, J. (1975) Biochem. Biophys. Res. Commun. 63, 1106-1112) has been extended from 27 residues to 39 residues. Also, it was found that two forms of Rhizopus pepsin differ in position 15, where Rhizopus pepsin I has an isoleucine and Rhizopus pepsin II a valine residue. The new sequences have been aligned with the amino-terminal sequences of penicillopepsin (EC 3.4.23.7), pig pepsin (EC 3.4.23.1), calf chymosin (EC 3.4.23.4), human pepsin (EC 3.4.23.2), human gastricsin (EC 3.4.23.3), and cow pepsin (EC 3.4.23.1). Residues 31-35 (numbering based on pig pepsin, Tang, J., Sepulveda, P., Marciniszyn, Jr., J., Chen, K.S.C., Huang, W.-Y. , Tao, N., Liu, D. & Lanier, P. (1973) Proc. Natl. Acad. Sci. U.S.A. 70, 3437-3739) are identical in all enzymes. This section contains one of the two aspartic acids (Asp-32) implicated in the active site. The similarity of the sequences provides strong evidence for the homology of these acid proteases.  相似文献   

10.
Transmembrane location of the retinal chromophore in the purple membrane of Halobacterium halobium was investigated in three different systems in which excitation energy transfer between the chromophore and external dye molecules condensed on the membrane surfaces was observed. In system ii, the energy donor was the retinal chromophore converted in situ to a fluorescent derivative. The fluorescent membranes were embedded in solid cobalt-EDTA, which served as energy acceptors. System iii was similar to system ii, except that the acceptors were tris(2,2′-bipyridyl)ruthenium(II) complex in solid form. The positively charged ruthenium complex had a radius of 0.7 nm, whereas the cobalt complex in system ii was smaller (radius ~0.4 nm) and negatively charged. System iv was stacked sheets of native purple membrane with interspersed ruthenium complex; energy transfer from the luminescent ruthenuim complex to the native retinal chromophore was observed. The energy transfer rates in these three systems, and in two additional systems already described (Kouyama, T., K. Kinosita, Jr., and A. Ikegami, 1983, J. Mol. Biol., 165:91-107), were all consistent with a location of the retinal chromophore at a depth of 1.0 ± 0.3 nm from a surface of the purple membrane. All the analyses in the present work involved an assumption that contacts between the external dye molecules and membrane surfaces were maximal; the depth values obtained cannot be underestimates. The chromophore therefore must be outside the middle one-third of the thickness, ~4.5 nm, of the purple membrane.  相似文献   

11.
By applying the same method used for F(1)-ATPase (TF(1)) from thermophilic Bacillus PS3 (Noji, H., Yasuda, R., Yoshida, M., and Kinosita, K., Jr. (1997) Nature 386, 299-302), we observed ATP-driven rotation of a fluorescent actin filament attached to the gamma subunit in Escherichia coli F(1)-ATPase. The torque value and the direction of the rotation were the same as those observed for TF(1). F(1)-ATPases seem to share common properties of rotation irrespective of the sources.  相似文献   

12.
The outer layer of the bacteriophage phi 6 nucleocapsid (NC) was removed by EDTA and reassociated with the core in the presence of Ca2+ or Mg2+. The core was relatively inaccessible to trypsin digestion, was composed of protein P1, and was in the dodecahedral framework reported previously. (H.T. Steely, Jr., and D. Lang, J. Virol. 51:479-483, 1984; Y. Yang and D. Lang, J. Virol. 51:484-488, 1984). The double-stranded RNA genome became RNase sensitive after EDTA treatment of the nucleocapsid.  相似文献   

13.
Arginase is a binuclear Mn(2+) metalloenzyme that catalyzes the hydrolysis of L-arginine to L-ornithine and urea. X-ray crystal structures of arginase complexed to substrate analogues N(omega)-hydroxy-L-arginine and N(omega)-hydroxy-nor-L-arginine, as well as the products L-ornithine and urea, complete a set of structural "snapshots" along the reaction coordinate of arginase catalysis when interpreted along with the X-ray crystal structure of the arginase-transition-state analogue complex described in Kim et al. [Kim, N. N., Cox, J. D., Baggio, R. F., Emig, F. A., Mistry, S., Harper, S. L., Speicher, D. W., Morris, Jr., S. M., Ash, D. E., Traish, A. M., and Christianson, D. W. (2001) Biochemistry 40, 2678-2688]. Taken together, these structures render important insight on the structural determinants of tight binding inhibitors. Furthermore, we demonstrate for the first time the structural mechanistic link between arginase and NO synthase through their respective complexes with N(omega)-hydroxy-L-arginine. That N(omega)-hydroxy-L-arginine is a catalytic intermediate for NO synthase and an inhibitor of arginase reflects the reciprocal metabolic relationship between these two critical enzymes of L-arginine catabolism.  相似文献   

14.
A solvation energy function for use in the molecular simulation of proteins is proposed. It is based on the accessible surface areas of atoms in the protein and on atomic solvation parameters derived from empirical vapor-to-water free energies of transfer of amino acid side-chain analogs. The energy function and its derivatives were added to the CHARMM molecular simulation program (Brooks, B.R., Bruccoleri, R.E., Olafson, B.D., States, D.J., Swaminathan, S., & Karplus, M., 1983, J. Comput. Chem. 4(2), 187-217). The effect of the added energy term was evaluated by 110 ps of molecular dynamics on the 26-residue protein melittin. The melittin monomer and tetramer were studied both with and without the added term. With the added energy term the monomer partially unfolded, while the secondary structure of the tetramer was preserved, in agreement with reported experiments (Brown, L.R., Lauterwein, J., & Wuethrich, K., 1980, Biochim. Biophys. Acta 622(2), 231-244; Lauterwein, J., Brown, L.R., & Wuethrich, K., 1980, Biochim. Biophys. Acta 622(2), 219-230).  相似文献   

15.
The interaction between Cu(I) superoxide dismutase and hydrogen peroxide   总被引:3,自引:0,他引:3  
The interaction between superoxide dismutase (SOD) and peroxide, under anaerobic conditions in the presence of an OH radical scavenger, formate, and an indicator, nitro blue tetrazolium, involves five reactions and an equilibrium: (table; see text) Reaction 3 occurs at a rate that is proportional to both peroxide and enzyme with no kinetic evidence for any intermediate peroxide-enzyme complex. Rate studies as a function of pH corroborate previously published work (Fuchs, H. J. R., and Borders, C. L., Jr. (1983) Biochem Biophys. Res. Commun. 116, 1107-1113; Blech, D. M., and Borders, C. L., Jr. (1983) Arch. Biochem. Biophys. 224, 579-586) suggesting that HO2-, and not H2O2, is the active species in this system: k(HO2- + superoxide dismutase-Cu+) = 2.6 x 10(3) M-1 s-1. Evidence is presented which suggests that HO2-, like O2-, reacts at rates that are affected by the electrostatic forces of the enzyme.  相似文献   

16.
We have investigated the reason for the sensitivity of the fluorescence excited-state lifetime of 1,6-diphenyl-1,3,5-hexatriene (DPH) and its phospholipid derivatives, 1-palmitoyl-2-[2-[4-(6-phenyl-trans-1,3,5- hexatrienyl)phenyl]ethyl)carbonyl)-3-sn-phosphatidylcholine (DPHpPC) and 1-palmitoyl-2-[2-[4-(6-phenyl-trans-1,3,5- hexatrienyl)phenyl]ethyl)carbonyl)-3-sn-phosphatidic acid (DPHpPA), to the concentration of these probes in dipalmitoylphosphatidylcholine (DPPC) multilamellar membranes (Barrow, D. A., and B. R. Lentz, 1985. Biophys. J. 48:221-234; Parente, R. A., and B. R. Lentz. 1985. Biochemistry. 24:6178-6185). We have interpreted self-quenching data, excitation and emission spectra, and phase and modulation lifetime data in terms of a model that envisions dimerization of these probes in a membrane bilayer. It is proposed that dimerization alters the symmetry of the DPH excited state so as to allow more rapid decay via the normally symmetry-disallowed route from the 1Ag* state. Global analysis of fluorescence phase shift and modulation ratio data for DPHpPC in terms of the dimerization model provided a good fit of these data as a function of both modulation frequency and probe concentration. Global analysis of a similar set of data for the charged phosphatide DPHpPA predicted that this probe was much less prone to dimerize than was the uncharged DPHpPC. This physically reasonable result provides support for the assumptions made in the development of our model. We conclude that the dimerization model allows rationalization of many of the anomalous photophysical properties of DPH and its derivatives in membranes.  相似文献   

17.
A M Schwartz  G D Fasman 《Biochemistry》1977,16(10):2287-2299
Complexes of DNA with polypeptides composed of Lys, Ala, and Gly in both a sequential order, poly(L-lysine-L-alanine-glycine), and a statistical distribution, poly(L-lysine36-L-alanine28-glycine), were prepared using gradient dialysis. These polypeptide-DNA complexes were studied using ultraviolet absorption (UV) and circular dichroism (CD) to probe the conformation, binding, and melting behavior of DNA in the complex. Complexes with the sequential polypeptide showed no structural change in the DNA; however, the complexes with the random polypeptide yield CD spectra similar to phi DNA [Maniatis, T., Venable, Jr., J.S., and Lerman, L.S. (1974), J. Mol. Biol. 84, 37]. A second sequential polypeptide, poly(L-Lys-L-Ala-L-Pro)n, -DNA complex was also studied. It was found to exhibit pronounced structural changes as a function of ionic strength and poly-peptide-DNA ratio, more similar to the random sequence that the ordered sequence of the Lys, Ala, Gly polymer. Thus the importance of the composition and amino acid sequence in polypeptides which bind to DNA, even in such simple systems, is demonstrated. Evidence from thermal denaturation, employing simultaneous monitoring of CD and UV changes, supports a model in which specific polypeptides cause condensation of the DNA in the complex into an asymmetric tertiary structure. The relevance of these model systems to chromatin is discussed.  相似文献   

18.
Low-light digitized video fluorescence microscopy has been utilized to measure the steady-state polarized fluorescence from the membrane probe diphenylhexatriene (DPH) and its cationic and phosphatidylcholine derivatives 1-(4-trimethylammoniumphenyl)-6-phenyl-1,3,5-hexatriene (TMA-DPH) and 2-[3-(diphenylhexatrienyl)propanoyl]-3-palmitoyl-L-alpha-phosphati dylcholine (DPH-PC), respectively, in cell-size (10-70 microns) unilamellar vesicles composed of gel-or fluid-phase phospholipid. Using an inverted microscope with epi-illumination optics and an intensified silicon intensified target camera interfaced to a minicomputer, fluorescence images of single vesicles were obtained at emission polarizer orientations of 0 degrees, 45 degrees, 90 degrees, and 135 degrees relative to the excitation light polarization direction. Fluorescence intensity ratios F90 degrees/F0 degrees (= F perpendicular/F parallel) and F135 degrees/F45 degrees were calculated on a pixel-by-pixel basis from digitized image pairs. Theoretical expressions were derived for collected polarized fluorescence as a function of position on the membrane surface as well as the degree of lipid order, in terms of the fluorophore's maximum angular motional freedom in the bilayer (identical to theta max), using a modification of the method of D. Axelrod (1979. Biophys. J. 26:557-574) together with the "wobbling-in-a-cone" model of probe rotational diffusion. Comparison of experimental polarization ratios with theoretical ratios yielded the following results. In gel-phase dipalmitoyl-phosphatidylcholine, the data for all three probes correspond to a model in which the cone angle theta max = 17 +/- 2 degrees and there exists a collective tilt of the phospholipid acyl chains of 30 degrees relative to the bilayer normal. In addition, approximately 5% of DPH and TMA-DPH molecules are aligned parallel to the plane of the bilayer. In fluid-phase palmitoyloleoyl-phosphatidylcholine, the data are well fit by models in which theta max = 60 +/- 2 degrees for DPH and DPH-PC and 32 +/- 4 degrees for TMA-DPH, with approximately 20% of DPH molecules and 10% of TMA-DPH molecules aligned parallel to the bilayer plane, and a net phospholipid tilt at or near the headgroup region of approximately 30 degrees. The results demonstrate that lipid order can be measured with a spatial resolution of approximately 1 micron2 in cell-size vesicles even with high aperture observation through a microscope.  相似文献   

19.
A biphasic response is seen at both 445 and 605 nm as the ascorbate—cytochrome c—cytochrome aa3 system is taken slowly from the anaerobic to the aerobic state. At low oxygen tensions the 445 nm band is more reduced while at high oxygen tensions the 605 nm band is more reduced. It is suggested that the redox potential for cytochrome a (contributing 70% at 605–630 nm and 50% at 445–455 nm) is a function of the redox state of cytochrome a3. This model can account for both the aerobic/anaerobic data and for observations of interactions in the anaerobic system alone (Leigh, Jr, J.S., Wilson, D.F., Owen, C.S. and King, T.E. (1974) Arch. Biochem. Biophys. 160, 476–486).  相似文献   

20.
The activity of rat liver glutaminase from sedimented fractions of freeze-thawed mitochondria is strongly affected by variation in pH over a physiologically relevant range at approximate physiological concentrations of activators. As pH increases from 7.1 to 7.7 at 0.7 mM ammonium and 10 mM phosphate, the S0.5 for glutamine decreases 3.5-fold, from 38 to 11 mM. This results in an 8-fold increase in reaction velocity at 10 mM glutamine. In addition, the M0.5 for phosphate activation decreases from 21 to 8.9 mM as pH increases from 7.1 to 7.7. This apparent effect of pH on the affinity of glutaminase for phosphate is similar to previous reports of the pH effect on activation by ammonium (Verhoeven, A. J., Van Iwaarden, J. F., Joseph, S. K., and Meijer, A. J. (1983) Eur. J. Biochem. 133, 241-244; McGivan, J. D., and Bradford, N. M. (1983) Biochim. Biophys. Acta 159, 296-302). Glutaminase does not respond to variation in pH between 7.1 and 7.7 when phosphate and ammonium are saturating. The effects of the two modifiers are additive. Each is still effective, as is pH, when the other is saturating. Therefore, it appears that the effects of pH on the apparent affinity of the enzyme for ammonium and phosphate account for the enzyme's response to pH. These results may help explain previous reports of minimal effects of pH on glutaminase at saturating concentrations of related substances (McGivan, J. D., Lacey, J. H., and Joseph, K. (1980) Biochim. J. 192, 537-542; Horowitz, M. L., and Knox, W. E. (1968) Enzymol. Biol. Clin. 9, 241-255; McGivan, J. D., and Bradford, N. M. (1983) Biochim. Biophys. Acta 759, 296-302). Glutaminase binds glutamine cooperatively with Hill coefficients ranging from 1.7 to 2.2, which suggests at least two and probably three or more interacting binding sites for glutamine. The strong response of liver glutaminase to pH and the fact that the reaction can supply metabolites for urea synthesis suggest a possible regulatory role of glutaminase in ureagenesis.  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