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.
The solubility properties of paramyosin in the zones of pH and ionic strength in which aggregation occurs were initially studied using preparations isolated by a method originally described by Bailey (Bailey, K. (1956), Pubbl. Stn. Zool. Napoli 29, 26). Other preparations yielding apparently different protein components have been described by Hodge (Hodge, A.J. (1952), Proc. Natl. Acad. Sci., U.S.A. 38, 850) using acid conditions, and Stafford and Yphantis (Stafford, W.F., AND Yphantis, D. (1972), Biochem. Biophys. Res. Commun. 49, 848) have identified alpha-, beta-, and gamma-paramyosin using various times and temperatures of extraction with or without ethylenediaminetetraacetic acid. We have found that acid-extracted paramyosin is very similar if not identical to alpha-paramyosin, but that both acid and alpha forms differ considerably from beta- and gamma-paramyosin. Beta-Paramyosin precipitates abruptly from solution in narrow zone of pH below neutrality, and increases in ionic strength shift the zone of precipitation toward lower pH values. In contrast, both acid and alpha-paramyosin show gradual aggregation with changing pH at lowerionic strength (less than 0.3) but sharp transitions similar to beta-paramyosin at higher ionic strength (greater than 0.3). Transitions were also found at lower pH (ca. 4.0) which were not mirror images of transitions at higher pH (ca. 7.0). Viscosity measurements show that acid extracted paramyosin is close in behavior to a native extract obtained by extraction in mild, nondenaturing media containing mixed antibiotics. Each of these extracts differed considerably from beta-paramyosin. Mild, nonhydrolytic procedures employed by others to remove small, noncovalent bonded components or to separate protein complexes were not effective in converting alpha- to beta-paramyosin. Comparison of extraction procedures strongly supports the suggestion of Stafford and Yphantis that beta- and gamma-paramyosin are hydrolytic products of alpha-paramyosin and that the proteases responsible may be of bacterial origin.  相似文献   

2.
R W Cowgill 《Biochemistry》1975,14(19):4277-4279
The helical muscle protein beta-paramyosin of 200,000 was treated by the general method of G. R. Jacobson et al. (1973), J. Biol. Chem. 248, 6583) for cleavage of the polypeptide chain at the site of Cys residues. The protein cleaved into two segments: CCF-1 of 140,000 daltons and CCF-2 of 60,000 daltons. The two segments were separated and some properties were compared. Circular dichroism measurements indicated that CCF-1 was completely helical and that CCF-2 was 85% in the alpha-helical form. The molecular size, resistance to pepsin digestion, stability to heat and urea, and solubility of CCF-1 were all similar to corresponding properties of a pepsin-resistant segment PPC-1 described earlier (Cowgill, R. W. (1972), Biochemistry 11, 4532). By contrast, the properties of CCF-2 were distinctly different. It was concluded that the CCF-1 segment, like the PPC-1 segment, arose from the N-terminal two-thirds of the paramyosin molecule. The CCF-2 segment from the C-terminal one-third of paramyosin had limited solubility at neutral pH that matched the low solubility of paramyosin. It was concluded that the CCF-2 region is responsible for the self-aggregating tendency of paramyosin at neutral pH and low ionic strength.  相似文献   

3.
The influence of ATP and KCl on the quaternary structure and the enzymatic activity of D-glyceraldehyde-3-phosphate dehydrogenase from yeast(Y-GAPDH) has been studied by ultracentrifugation, gel chromatography and standard optical tests. In 0.1 M imidazole buffer pH 7.0, at low temperature (0°C) both complete deactivation and dissociation to dimers occur in the presence of 2 mM ATP and 0.1 M 2-mercaptoethanol. In 0.067 M phosphate buffer pH 7.0, containing 2 mM ATP and 1 mM dithiothreitol, only slight deactivation paralleled by minor changes of the native quaternary structure is observed. In this same buffer, increasing temperature leads to stabilization of both the tetrameric state and the catalytic activity of the enzyme. Deactivation and dissociation in the presence of 0.15 M KCl (in 0.2 M glycine buffer 9.1 ≥ pH ≥ 8.0) is a function of pH rather than electrolyte concentration; at neutral pH the enzyme is stabilized in its native state. Contrary to earlier assumptions in the literature, ATP and KCl under the above experimental conditions do not appear to play an important role in the in vivo regulation of Y-GAPDH.  相似文献   

4.
A calorimetric titration method was used to study ADP binding to native myosin. Data were analyzed by assuming that the myosin molecule has n independent and identical sites for ADP binding. The enthalpy change (deltaH), the binding constant (K), and n were determined. In 0.5 M KCl, 0.01 M MgCl2, and 0.02 M Tris/HCl (pH 7.8), we found: at 0 degrees, deltaH = -57.1 +/- 3.2 kJ-mol-1, log K = 6.42 +/- 0.13, n = 1.49 +/- 0.07; at 12 degrees, deltaH = 73.1 +/- 3.2 kJ-mole-1, log K = 6.08 +/- 0.13, and n = 1.74 +/- 0.07. The average heat capacity change on ADP binding to myosin between 0 and 12 degrees is thus -1.4 +/- 0.4 kJ-mol-1-K-1. Reasonably consistent results were obtained at 25 degrees, suggesting ADP binding to myosin is as strongly exothermic as at lower temperatures, although further interpretation of this result seems unwarranted, mainly because of the instability of myosic at this temperature. The number of protons released on binding of ADP to myosin was determined in separate experiments. The value was 0.19 +/- 0.02 at both 0 and 12 degrees. The reaction of protons with Tris thus contributes about -9.5 kJ-mol-1 to the observed heat on ADP binding.  相似文献   

5.
The terminal web of the intestinal brush border contains a spectrin- like protein, TW 260/240 (Glenney, J. R., Jr., P. Glenney, M. Osborne, and K. Weber, 1982, Cell, 28:843-854.) that interconnects the "rootlet" ends of microvillar filament bundles in the terminal web (Hirokawa, N., R. E. Cheng, and M. Willard, 1983, Cell, 32:953-965; Glenney J. R., P. Glenney, and K. Weber, 1983, J. Cell Biol., 96:1491-1496). We have investigated further the structural properties of TW 260/240 and the interaction of this protein with actin. Salt extraction of TW 260/240 from isolated brush borders results in a loss of terminal web cross- linkers primarily from the apical zone directly beneath the plasma membrane. Morphological studies on purified TW 260/240 using the rotary shadowing technique confirm earlier results that this protein is spectrin-like and is in the tetrameric state in buffers of low ionic strength. However, examination of TW 260/240 tetramers by negative staining revealed a molecule much straighter and more uniform in diameter than rotary-shadowed molecules. At salt concentrations at (150 mM KCl) and above (300 mM KCl) the physiological range, we observed a partial dissociation of tetramers into dimers that occurred at both 0 degree and 37 degrees C. We also observed (in the presence of 75 mM KCl) a concentration-dependent self-association of TW 260/240 into sedimentable aggregates. We have studied the interaction of TW 260/240 with actin using techniques of co-sedimentation, viscometry, and both light and electron microscopy. We observed that TW 260/240 can bind and cross-link actin filaments and that this interaction is salt- and pH- dependent. Under optimum conditions (25-75 mM KCl, at pH 7.0) TW 260/240 cross-linked F-actin into long, large-diameter bundles. The filaments within these bundles were tightly packed but loosely ordered. At higher pH (7.5) such bundles were not observed, although binding and cross-linking were detectable by co-sedimentation and viscometry. At higher salt (greater than 150 mM KCl), the binding of TW 260/240 to actin was inhibited. The presence of skeletal muscle tropomyosin had no significant effect on the salt-dependent binding of TW 260/240 to F- actin.  相似文献   

6.
Exchange assays have been validated to study several forms of the progesterone receptor found to occur in nuclei of rat placenta after extraction with high salt. One form was solubilized by the extraction procedure (KCl extractable Rpn) and another form remained attached to nuclear structures (KCl resistant Rpn). Specific binding of progesterone was optimized in both forms using buffered media containing 0.01 M Tris, 30%-glycerol (v/v), 0.2 mM leupeptin, and 1 mM dithiothreitol (TDGL), pH 7.8, at 0-4 degrees C for 18-24 h. At 0-4 degrees C the nuclear receptors were stable and degradation was negligible even after 44 h of in vitro incubation. The binding reaction between progesterone and receptor demonstrated mass action principles of ligand exchange throughout this interval. Saturation analysis indicated the presence of a single binding moiety of high affinity (app Kd = 2.9-3.2 nM) for both forms of the receptor. However, the nuclear progesterone receptor was thermolabile and after a 10 min exposure to 30 degrees C no longer complexed ligand. At an intermediate incubation temperature of 22 degrees C the binding reaction was stable for about 30 min. The KCl resistant binding sites were markedly more thermolabile. Addition of 10 mM Na molybdate protected all forms of the nuclear progesterone receptor from thermal denaturation and extended the life of the complex 3-4-fold. The dissociation rate constant of progesterone-nuclear receptor complex in each preparation was 6-8 X 10(5) s-1 resulting in a half-life of about 3 h. The KCl resistant and extractable binding sites were sensitive to blockade by 1 mM N-ethylmaleimide which was reversed by co-incubation with a 2-fold molar excess of dithiothreitol. This suggested that reduced sulfhydryl groups located on or near the surface of the ligand binding domain of the receptor were necessary to bind hormone. These studies showed that the interactions between ligand and the KCl resistant and extractable receptor sites found in rat placenta were of high affinity, saturable, and heat sensitive. Thus, these binding moieties exhibited physicochemical behavior very similar to each other and to the placental receptor which has previously been partially purified from the cytosol. The conclusion is made that all of the nuclear receptor binding sites for progesterone are structurally identical. Thus, the distinctive physicochemical properties responsible for KCl resistant and extractable forms of the nuclear progesterone receptor must reside in other domains of the receptor molecule.  相似文献   

7.
Some biophysical properties of a (Na+, K+)-ATPase preparation from guinea-pig kidney have been analysed. The recently developed technique of laser Doppler spectroscopy was applied to measure particle mobility under electrophoretic conditions. The following results were obtained: 1. magnesium ions at pH 7.3 decrease the mobility of the ATPase containing vesicles by binding to negatively charged surface groups. At pH 3.3 the competitive binding of protons causes a shift of the mobility vs. [Mg2+] curve to higher values of [Mg2+], 2. binding of ATP at pH 7.3 (Kd = 0.9 X 10(-4) M for (mM 1 NaCl, 0.2 KCl, 0.1 MgCl2, 0.1 Tris) was measured as an increase in particle mobility depending also on [Mg2+]. At pH 3.3 also unspecific ATP-binding occurred, 3. ITP and GTP had the same Kd value as ATP; ADP a slightly lower one (Kd = 1.2 X 10(-4) M). Tris-H3PO4 (Kd = 2.6 X 10(-4) M) was also able to increase particle mobility, but only at higher concentrations and not to the same extent as ATP; AMP induced only very small changes, 4. from the mobility-pH curve an isoelectric point of 4.1 is derived (buffer: 1 mM NaCl, 0.2 mM KCl, 0.1 mM MgCl2, 0.1 mM Tris). In the presence of 0.9 mM ATP the isoelectric point is shifted to 3.2. As the electrophoretic mobility is directly proportional to the net charge of the vesicles, the results may be interpreted as changes in surface charge density, originating from both a conformational change of the ATPase polypeptide and a decrease in vesicle size.  相似文献   

8.
T Tao  J Xie  M L Drumm  J Zhao  P B Davis    J Ma 《Biophysical journal》1996,70(2):743-753
The cystic fibrosis transmembrane conductance regulator (CFTR) chloride channel exhibits multiple subconductance states. To study the regulation of conductance states of the CFTR channel, we expressed the wild-type CFTR protein in HEK 293 cells, and isolated microsomal membrane vesicles for reconstitution studies in lipid bilayer membranes. A single CFTR channel had a dominant conductance of 7.8 pS (H), plus two sub-open states with conductances of approximately 6 pS (M) and 2.7 pS (L) in 200 mM KCl with 1 mM MgCl2 (intracellular) and 50 mM KCl with no MgCl2 (extracellular), with pH maintained at 7.4 by 10 mM HEPES-Tris on both sides of the channel. In 200 mM KCl, both H and L states could be measured in stable single-channel recordings, whereas M could not. Spontaneous transitions between H and L were slow; it took 4.5 min for L-->H, and 3.2 min for H-->L. These slow conversions among subconductance states of the CFTR channel were affected by extracellular Mg; in the presence of millimolar Mg, the channel remained stable in the H state. Similar phenomena were also observed with endogenous CFTR channels in T84 cells. In high-salt conditions (1.5 M KCl), all three conductance states of the expressed CFTR channel, 12.1 pS, 8.2 pS, and 3.6 pS, became stable and seemed to gate independently from each other. The existence of multiple stable conductance states associated with the CFTR channel suggests two possibilities: either a single CFTR molecule can exist in multiple configurations with different conductance values, or the CFTR channel may contain multimers of the 170-kDa CFTR protein, and different conductance states are due to different aggregation states of the CFTR protein.  相似文献   

9.
Paramyosin was prepared from earthworms (Lumbricus terrestris) by two different methods that have been used in the past. Polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate shows that the older method yields slightly degraded material (mostly β- and γ-paramyosin) while the newer method yields essentially intact, i.e., α-, paramyosin. Physical studies, particularly circular dichroism, light scattering, and sedimentation velocity show that the native molecule is a double α-helical coiled coil of molecular weight 200,000, length 1200 Å, and diameter 20 Å. These properties are the same as reported previously for molluscan paramyosin. Also like clam paramyosin, the worm protein molecule loses its helix content and dissociates into its two constituent polypeptide chains upon exposure to sufficient concentration of Gdn-HCl. Furthermore, the same partially denatured states can be reached from either native or completely denatured proteins, indicating that they are all equilibrium states. However, the Gdn · HCl-induced denaturation profile for the worm paramyosin is quite different from the clam. The helix content of worm paramyosin diminishes monophasically with increasing concentration of Gdn-HCl, showing that the molecule does not possess a region of special stability such as its clam analog boasts. This conclusion is supported by experiments on papain digestion of worm paramyosin, wherein no resistent core is seen.  相似文献   

10.
The Na(+)-ATPase activity of Na+,K(+)-ATPase in the absence of K+ was least dependent on the sodium concentration when the pH was 9.5. Around 40% of the phosphoenzyme formed from ATP in the presence of 0.5 mM MgCl2 at alkaline pH was insensitive to both KCl and ADP. High-Na+ chase reversed this insensitivity, i.e., the phosphoenzyme became sensitive to KCl or ADP. On the other hand, phosphorylation at 0.1 mM MgCl2 instead of 0.5 mM showed at least 95% sensitivity to KCl. These observations suggest that ADP- and KCl-insensitive phosphoenzyme was formed when excess Mg++ was present during phosphorylation at alkaline pH. This phosphoenzyme might be an intermediate in the process of ATP hydrolysis.  相似文献   

11.
o-Diphenol oxidase activities (o-diPO) of chemically modified functional unit RvH1-a of molluscan hemocyanin Rapana venosa were studied using L-Dopa and dopamine as substrates. With L-Dopa as substrate the native FU RvH1-a did not show any o-diPO activity. Therefore the native FU RvH1-a was converted to enzymatic active form, after treatment with SDS, trypsin, urea and different values of pH when its o-diPO activity was studied. The highest artificial induction of o-diPO activity was observed after incubation of FU with 3.0mM SDS, and RvH1-a shows both, dopamine (K(M)=6.53mM, k(cat)/K(M)=1.29) and L-Dopa (K(M)=2.0mM, k(cat)/K(M)=2.1) activity due to a more open active site of the enzyme and better access of the substrates. It was determined that the K(M) value of SDS-activated RvH1-a against dopamine is higher compared to those of hemocyanins from Helix vulgaris, Helix pomatia and native tyrosinase from Ipomoea batatas but much lower than that from Illex argentinus (ST94) tyrosinase and arthropodan hemocyanin from Carcinus aestuarii. The Km value of SDS-activated RvH1-a against L-Dopa is higher than those of hemocyanins from H. vulgaris and Cancer magister, but lower than that of the tyrosinase from Streptomyces albus.  相似文献   

12.
M Y Lee  C K Tan  K M Downey  A G So 《Biochemistry》1984,23(9):1906-1913
DNA polymerase delta from calf thymus has been purified to apparent homogeneity by a new procedure which utilizes hydrophobic interaction chromatography with phenyl-Sepharose at an early step to separate most of the calcium-dependent protease activity from DNA polymerase delta and alpha. The purified enzyme migrates as a single protein band on polyacrylamide gel electrophoresis under nondenaturing conditions. The sedimentation coefficient of the enzyme is 7.9 S, and the Stokes radius is 53 A. A molecular weight of 173K has been calculated for the native enzyme. Sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the homogeneous enzyme reveals two polypeptides of 125 and 48 kDa. This subunit structure differs from that of DNA polymerase delta prepared by our previous procedure, which was composed of subunits of 60 and 49 kDa [Lee, M. Y. W. T., Tan, C.-K., Downey , K. M., & So, A. G. (1981) Prog . Nucleic Acid Res. Mol. Biol. 26, 83-96], suggesting that the 60-kDa polypeptide may have been derived from the 125-kDa polypeptide during enzyme purification, possibly as the result of cleavage of an unusually sensitive peptide bond. DNA polymerase delta is separated from DNA polymerase alpha by hydrophobic interaction chromatography on phenyl-Sepharose; DNA polymerase delta is eluted at pH 7.2 and DNA polymerase alpha at pH 8.5. DNA polymerase delta can also be separated from DNA polymerase alpha by chromatography on hydroxylapatite; DNA polymerase alpha binds to hydroxylapatite in the presence of 0.5 M KCl, whereas DNA polymerase delta is eluted at 90 mM KCl.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

13.
Arrangement of myosin heads on Limulus thick filaments   总被引:3,自引:2,他引:1       下载免费PDF全文
The two myosin heads with a single surface subunit on thick filaments from chelicerate arthropod muscle may originate from the same, or from axially sequential molecules, as suggested by three-dimensional reconstructions. The resolution attained in the reconstructions, however, does not permit one to distinguish unequivocally between these two possible arrangements. We examined the effect of 0.6 M KCl on relaxed thick filaments separated from Limulus muscle and filaments in which nearest myosin heads were cross-linked by the bifunctional agent, 3,3'-dithio-bis[3'(2')-O-[6-propionylamino)hexanoyl]adenosine 5'-triphosphate (bis22ATP), in the presence of vanadate (Vi). In high salt, surface myosin dissolved from both native, relaxed filaments and those exposed to 1-2 mM dithiothreitol after cross-linking, but was retained on filaments with cross-linked heads. Since bis22ATP must form intermolecular bonds between myosin heads within each subunit to prevent myosin solubilization in high salt, we conclude that each of these heads originates from a different myosin molecule, as was previously predicted by the reconstructions.  相似文献   

14.
Na+, K+-dependent ATPase [EC 3.6.1.3] was purified from porcine kidney by the method of Lane et al. [(1973) J. Biol. Chem. 248, 7197-7200] with slight modifications [Yamaguchi, M. & Tonomura, Y., (1979) J. Biochem. 86, 509-523]. The amounts of a phosphorylated intermediate (EP) and ouabain bound to the enzyme during the ATPase reaction were measured in 2.1 mM MgCl2 and various concentrations of NaCl and KCl at pH 7.5 and 20 degrees C. In presence of NaCl and the absence of KCl, the molar ratio of the amounts of EP and bound ouabain was 1 : 2. In the presence of both NaCl and KCl, it was 1 : 1. In both cases, the amount of bound ouabain was equal to that of EP in the absence of ouabain. These findings suggest that the functional unit of the transport ATPase is a dimer.  相似文献   

15.
Counting of integral numbers of cysteine residues of the reduced and denaturated form of cyclomaltodextrin glucanotransferase (CGTase) from Bacillus circulans var. alkalophilus (ATCC 21783) showed two cysteine residues per enzyme molecule. Titrations of the enzyme with 5,5'-dithiobis-(2-nitrobenzoic acid) led to the same result. No free SH-group was detected in denatured form of CGTase, indicating that the two cysteine residues are linked by one disulfide bridge. Cyclizing activity of the GdmCl-denaturated and reduced enzyme was 13% of that of the native one. Incubation of CGTase with diethylpyrocarbonate (DEP) showed a pseudo-first-order inhibition with second-order rate constant of 3.2 M-1 s-1. Reaction with hydroxylamine and spectroscopic studies implied that inactivation of CGTase by DEP is due to modification of one histidine residue concomitantly with a 50% decrease in the cyclizing activity (t1/2 = 10.8 min). The inhibition was partially reversible. CGTase was protected against inactivation by alpha- and beta-cyclodextrins suggesting that the modified histidine residue is at or near the active site. Conversion of starch with DEP-modified enzyme resulted in a decreased formation of cyclodextrins while the relative amount of reducing sugars increased. Preliminary results on modification of CGTase with other reagents, e.g., Woodward's reagent K, 2,3-butanedione and carbodiimide are included.  相似文献   

16.
The "hypothetical protein" AQ_1862 was isolated from the membrane fraction of Aquifex aeolicus and identified as the major porin. In experiments with one conducting unit (molecule) a conductance of 1.4 nS was observed in 0.1 M KCl at pH 7.5. This stable (basic) conductance was superimposed by conductance fluctuations of approximately 0.25 nS. Because both events were always observed simultaneously, it is suggested that they are caused by the same molecular entity. Nonetheless they show very different properties. The basic conductance is anion selective at neutral pH with a conductance sequence Cl- approximately Br- approximately NO3->F->gluconate approximately acetate approximately propionate and does not saturate up to 0.5 M KCl. At alkaline pH and in the presence of large anions, it becomes unselective and the conductance saturates at low concentrations (Km approximately 20 mM). In contrast the fluctuating component is mainly cation selective with a conductance sequence K+ approximately Rb+>NH4+>Na+ approximately Li+ approximately Cs+. It saturates at low salt concentrations (Km approximately 15 mM) and is not affected by pH. In view of the diverging properties of both conductance components, it seems appropriate to assume that AQ_1862 has two different conducting pathways rather than one with two different open states.  相似文献   

17.
Klebsiella aerogenes urease was purified 1,070-fold with a 25% yield by a simple procedure involving DEAE-Sepharose, phenyl-Sepharose, Mono Q, and Superose 6 chromatographies. The enzyme preparation was comprised of three polypeptides with estimated Mr = 72,000, 11,000, and 9,000 in a alpha 2 beta 4 gamma 4 quaternary structure. The three components remained associated during native gel electrophoresis, Mono Q chromatography, and Superose 6 chromatography despite the presence of thiols, glycols, detergents, and varied buffer conditions. The apparent compositional complexity of K. aerogenes urease contrasts with the simple well-characterized homohexameric structure for jack bean urease (Dixon, N. E., Hinds, J. A., Fihelly, A. K., Gazzola, C., Winzor, D. J., Blakeley, R. L., and Zerner, B. (1980) Can. J. Biochem. 58, 1323-1334); however, heteromeric subunit compositions were also observed for the enzymes from Proteus mirabilis, Sporosarcina ureae, and Selemonomas ruminantium. K. aerogenes urease exhibited a Km for urea of 2.8 +/- 0.6 mM and a Vmax of 2,800 +/- 200 mumol of urea min-1 mg-1 at 37 degrees C in 25 mM N-2-hydroxyethylpiperazineN'-2-ethanesulfonic acid, 5.0 mM EDTA buffer, pH 7.75. The enzyme activity was stable in 1% sodium dodecyl sulfate, 5% Triton X-100, 1 M KCl, and over a pH range from 5 to 10.5, with maximum activity observed at pH 7.75. Two active site groups were defined by their pKa values of 6.55 and 8.85. The amino acid composition of K. aerogenes urease more closely resembled that for the enzyme from Brevibacter ammoniagenes (Nakano, H., Takenishi, S., and Watanabe, Y. (1984) Agric. Biol. Chem. 48, 1495-1502) than those for plant ureases. Atomic absorption analysis was used to establish the presence of 2.1 +/- 0.3 mol of nickel per mol of 72,000-dalton subunit in K. aerogenes urease.  相似文献   

18.
An intermediate-conductance K+ channel (I.K.), the activity of which is increased by hyperpolarization, was previously identified in the lateral membrane of the cortical collecting duct (CCD) of the rat kidney (Wang, W. H., C. M. McNicholas, A. S. Segal, and G. Giebisch. 1994. American Journal of Physiology. 266:F813-F822). The biophysical properties and regulatory mechanisms of this K+ channel have been further investigated with patch clamp techniques in the present study. The slope conductance of the channel in inside-out patches was 50 pS with 140 mM KCl in the pipette and 5 mM KCl, 140 mM NaCl (NaCl Ringer''s solution) in the bath. Replacement of the bath solution with symmetrical 140 mM KCl solution changed the slope conductance of the channel to 85 pS and shifted the reversal potential by 55 mV, indicating that the selectivity ratio of K+/Na+ was at least 10:1. Channel open probability (Po) in inside-out patches was 0.12 at 0 mV and was increased by hyperpolarization. The voltage-dependent Po was fitted with the Boltzmann''s equation: Po = 1/[1 + exp(V-V1/2)zF/RT], with z = 1.2 and V1/2 = -40 mV. Addition of 2 mM tetraethylammonium or 500 mM quinidine to the bath blocked the activity of the K+ channel in inside-out patches. In addition, decrease in the bath pH from 7.40 to 6.70 reduced Po by 30%. Addition of the catalytic subunit of protein kinase A (PKAc; 20 U/ml) and 100 microM [corrected] MgATP to the bath increased Po from 0.12 to 0.49 at 0 mV and shifted the voltage dependence curve of channel activity toward more positive potentials by 40 mV. Two exponentials were required to fit both the open-time and the closed-time histograms. Addition of PKAc increased the long open-time constant and shortened the long closed-time constant. In conclusion, PKA-mediated phosphorylation plays an important role in the regulation of the voltage dependence of the hyperpolarization-activated K+ channel in the basolateral membrane of CCD.  相似文献   

19.
Phosphorylation of the thick filament proteins myosin and paramyosin was studied in Caenorhabditis elegans. We have incubated partially purified, native thick filaments with [gamma 32P] ATP in the presence of 50-750 mM NaCl, pH 6.5-8.0. Myosin heavy chain and paramyosin were phosphorylatable only upon solubilization at 450 mM and higher NaCl concentrations. Under conditions preserving native structures, no phosphorylation of these proteins occurred. The phosphorylation required Mg2+ but was unaffected by cAMP, cGMP or Ca2+. The specific inhibitor of cAMP and cGMP kinase catalytic subunits, H8, inhibits the activity. Sedimentation experiments show that the kinase may associate with but is not an intrinsic component of thick filaments. In C. elegans, phosphorylation by the thick filament associated activity of myosin and paramyosin is dependent upon the state of their assembly.  相似文献   

20.
Competition experiments using 9-anthroylcholine, a fluorescent dye that undergoes calmodulin-dependent binding by smooth muscle myosin light chain kinase [Malencik, D. A., Anderson, S. R., Bohnert, J. L., & Shalitin, Y. S. (1982) Biochemistry 21, 4031], demonstrate a strongly stabilizing interaction between the adenosine 5'-triphosphate and myosin light chain binding sites operating within the enzyme-calmodulin complex but probably not in the free enzyme. The interactions in the latter case may be even slightly destabilizing. The fluorescence enhancement in solutions containing 5.0 microM each of the enzyme and calmodulin is directly proportional to the maximum possible concentration of bound calcium on the basis of four calcium binding sites. Evidently, all four calcium binding sites of calmodulin contribute about equally to the enhanced binding of 9-anthroylcholine by the enzyme. Fluorescence titrations on solutions containing 1.0 microM enzyme plus calmodulin yield a Hill coefficient of 1.2 and K = 0.35 +/- 0.08 microM calcium. Three proteolytic fragments of smooth muscle myosin light chain kinase, apparent products of endogenous proteolysis, were isolated and characterized. All three possess calmodulin-dependent catalytic activity. Their interactions with 9-anthroylcholine, in both the presence and absence of calmodulin, are similar to those of the native enzyme. However, the stabilities of their complexes with calmodulin vary. The corresponding dissociation constants range from 2.8 nM for the native enzyme and 8.5 nM for the 96K fragment to approximately 15 nM for the 68K and 90K fragments [0.20 N KCl, 50 mM 3-(N-morpholino)propanesulfonic acid, and 1 mM CaCl2, pH 7.3, 25 degrees C]. A coupled fluorometric assay, modified from a spectrophotometric assay for adenosine cyclic 3',5'-phosphate dependent protein kinase [Cook, P. F., Neville, M. E., Vrana, K. E., Hartl, F. T., & Roskoski, R. (1982) Biochemistry 21, 5794], has provided the first continuous recordings of myosin light chain kinase phosphotransferase activity. The results show that smooth muscle myosin light chain kinase is a responsive enzyme, whose activity adjusts rapidly to changes in solution conditions.  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