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.
Estrogen-induced autocrine and paracrine growth factors are thought to stimulate endometrial proliferation. However, the proliferation is arrested at an early secretory phase although the amount of growth factors and their receptors remains constant. These receptors are protein tyrosine kinases which cause activating receptor autophosphorylation and phosphorylation of signalling substances. One inhibitory mechanism is the reverse dephosphorylation by phosphatases hydrolysing phosphotyrosines. Previously, an acid phosphotyrosine phosphatase activity was found in endometrial secretory glands. The purpose of this study was to evaluate its characteristics. Catalytic and immunohistochemical techniques were applied on sections obtained from human endometrium and other tissues. Endometrial acid phosphatase hydrolysed phosphotyrosine, not only at acid, but also at neutral pH values. An alternative substrate was α-naphthyl phosphate or β-glycerophosphate but not phosphoserine. Activities were inhibited by tartrate and fluoride but not by formaldehyde. These catalytic properties are identical only to those of prostatic acid phosphatase (PAP). A PAP-like nature was also proved by positive PAP immunohistochemistry. In conclusion, endometrial glands contain a phosphotyrosine phosphatase which is identical to PAP. Its activity is menstrual-cycle-dependent, being present only at the secretory phase, and it may counterbalance receptor tyrosine kinases terminating glandular proliferation despite constant levels of growth factors and their receptors.  相似文献   

2.
The phosphotyrosyl protein phosphatase activity of prostatic acid phosphatase (PAP) has been well established. It has also been suggested that PAP partly regulates the activity of growth factor receptors by dephosphorylating the autophosphorylysable tyrosines in them. We studied the binding of the peptides from epidermal growth factor receptor (EGFR) and its homolog (ErbB-2), corresponding to their autophosphorylation sites, to PAP using theoretical modeling and molecular dynamics (MD) simulation methods. Nine different peptides, each with a phosphotyrosine residue, were docked on human PAP. The binding energies of these peptide-PAP complexes were calculated theoretically and compared to experimentally obtained affinities. The peptide Ace--DNLpYYWD--NH2 from ErbB-2(1197-1203) showed the most favorable free energy of binding when estimated theoretically. The results demonstrate that the presence of another tyrosine residue proximate to C-terminal of autophosphorylysable Tyr enhances the binding affinity considerably. The presence of a bulky group instead prevents the binding, as is observed in case of peptide Ace--NLYpYWDQ--NH2 which failed to bind, both in theoretical calculations and experiments. Thus we demonstarted that PAP could potentially bind to EGFR and Erbb-2 and dephosphorylate them. Thus it could be involved in the regulation of the function of such receptors. In addition, complexes of a peptide from AngiotensinII and phosphotyrosine(pY) with human PAP were also modeled. The effects of different protonation states of the titratable active site residues on ligand (pY) binding have also been investigated. For a favorable binding His12 and Asp258 should be neutral, His257 should be positively charged and the phosphate group of the ligand should be in PO(4) (3-) state. Furthermore, the analysis of protein motion as observed during simulations suggests the loop-loop contact in the PAP dimer to be of importance in cooperativity.  相似文献   

3.
Some properties of hexameric purine nucleoside phosphorylase II (EC 2.4.2.1) from Escherichia coli K-12 were studied. The enzyme obeys the Michaelis-Menten kinetics with respect to purine substrates (Km for inosine, deoxyinosine and hypoxanthine are equal to 492, 106 and 26.6 microM, respectively) and exhibits negative kinetic cooperativity towards phosphate and ribose-1-phosphate. The Hill coefficient is equal to approximately 0.5 for both substrates. Hexameric purine nucleoside phosphorylase II is not a metal-dependent enzyme; its activity is inhibited by Cu2+, Zn2+, Ni2+ and SO4(2-). The enzyme is the most stable at pH 6.0; it contains essential thiol groups. All substrates partly protect the enzyme against inactivation by 5.5'-dithiobis(2-nitrobenzoic acid) and heat-inactivation and, with the exception of phosphate-against inactivation by p-chloromercuribenzoate. Hypoxanthine, especially in combination with phosphate, afford the best protection against inactivation.  相似文献   

4.
A partially purified bovine cortical bone acid phosphatase, which shared similar characteristics with a class of acid phosphatase known as tartrate-resistant acid phosphatase, was found to dephosphorylate phosphotyrosine and phosphotyrosyl proteins, with little activity toward other phosphoamino acids or phosphoseryl histones. The pH optimum was about 5.5 with p-nitrophenyl phosphate as substrate but was about 6.0 with phosphotyrosine and about 7.0 with phosphotyrosyl histones. The apparent Km values for phosphotyrosyl histones (at pH 7.0) and phosphotyrosine (at pH 5.5) were about 300 nM phosphate group and 0.6 mM, respectively, The p-nitrophenyl phosphatase, phosphotyrosine phosphatase, and phosphotyrosyl protein phosphatase activities appear to be a single protein since these activities could not be separated by Sephacryl S-200, CM-Sepharose, or cellulose phosphate chromatographies, he ratio of these activities remained relatively constant throughout the purification procedure, each of these activities exhibited similar thermal stabilities and similar sensitivities to various effectors, and phosphotyrosine and p-nitrophenyl phosphate appeared to be alternative substrates for the acid phosphatase. Skeletal alkaline phosphatase was also capable of dephosphorylating phosphotyrosyl histones at pH 7.0, but the activity of that enzyme was about 20 times greater at pH 9.0 than at pH 7.0. Furthermore, the affinity of skeletal alkaline phosphatase for phosphotyrosyl proteins was low (estimated to be 0.2-0.4 mM), and its protein phosphatase activity was not specific for phosphotyrosyl proteins, since it also dephosphorylated phosphoseryl histones. In summary, these data suggested that skeletal acid phosphatase, rather than skeletal alkaline phosphatase, may act as phosphotyrosyl protein phosphatase under physiologically relevant conditions.  相似文献   

5.
Sopina VA 《Tsitologiia》2001,43(7):701-707
Activity and thermoresistance of acid phosphatase were determined in supernatant of Amoeba proteus homogenates using 1-naphthyl phosphate (pH 4.0) and p-nitrophenyl phosphate (pH 5.5). Although tartrate-resistant and tartrate-sensitive acid phosphatases hydrolyse both substrates, the former mainly hydrolyses p-nitrophenyl phosphate and the latter 1-naphthyl phosphate. A decrease in the activity of the total and tartrate-sensitive acid phosphatases, when using 1-naphthyl phosphate, and of the total and tartrate-resistant acid phosphatases, when using p-nitrophenyl phosphate, was found in amoebae acclimated to 10 degrees C (10 degrees-amoebae) compared to those acclimated to 25 degrees C (25 degrees-amoebae). Using 1-naphthyl phosphate, the thermoresistance of the total acid phosphatase was lower in 10 degrees-amoebae than in 25 degrees-amoebae, but the thermostability of tartrate-resistant enzyme was the same in both groups of amoebae. Using p-nitrophenyl phosphate, the thermoresistance of the total and tartrate-resistant acid phosphatases was lower (the latter only slightly) in 10 degrees-amoebae than in 25 degrees-amoebae. It is suggested that at least with the use of 1-naphthyl phosphate a decrease in thermostability of the total acid phosphatase may be due to a decrease in thermoresistance of tartrate-sensitive enzyme. The results obtained confirm the author's previous data on the activity and thermostability of electrophoretic forms of acid phosphatase using 2-naphthyl phosphate in 10- and 25 degrees-amoebae (Sopina, 2001). It is the first case of discovering a correlation between changes in primary cell thermoresistance of amoebae cultured at different temperatures and changes in the activity and thermostability of acid phosphatase in their homogenates, with the number of electrophoretic forms of this enzyme and their mobility being permanent.  相似文献   

6.
Uric acid inhibited 50% of the activity of bovine kidney low molecular mass phosphotyrosine protein phosphatase at concentrations of 1.0, 0.4, 1.3, and 0.2 mM, respectively for p -nitrophenyl phosphate (p -NPP), flavine mononucleotide, β -naphthyl phosphate and tyrosine phosphate (Tyr-P) as substrates. The mixed type inhibition of p -NPP hydrolysis was fully reversible, with K ic and K iu values of 0.4 and 1.1 mM, respectively; the inhibition by uric acid shifted the pH optimum from 5.0 to 6.5. When Tyr-P was the substrate, competitive inhibition was observed with a K i value of 0.05 mM. Inhibition studies by uric acid in the presence of thiol compounds, and preincubation studies in the presence of inorganic phosphate suggest that the interaction of uric acid with the enzyme occurred at the active site, but did not involve SH residues, and that the mechanism of inhibition depended on the structure of the substrates.  相似文献   

7.
The sensitivity (change of flux per unit change in the concentration of substrate) and response (change of flux per unit change in the concentration of modifier) are studied for a two-site Adair model in which cooperativity arises from both binding and catalytic interactions. For positive cooperativity, the sensitivity is weakly dependent on the Hill coefficient for the binding case, but can increase without limit for the catalytic case. Negatively cooperative enzymes (binding only) give very large sensitivities compared with positively or non-interacting systems, but the sensitivity rapidly decreases as the saturation increases above 25%. Modifiers greatly enhance the sensitivity; large changes in flux can be obtained for small changes in the concentrations of substrates and modifiers. In general, increasing the degree of kinetic cooperativity decreases the degree of binding cooperativity; selective pressure to maximize the sensitivity and response of allosteric enzymes may act to optimize cooperativity of binding modifiers and kinetic cooperativity of substrate turnover. The initial velocity equations including modifiers can be extended to bi-substrate, cooperative kinetics. The kinetics of methanol dehydrogenase are discussed.  相似文献   

8.
Kinetic properties of rat liver acid phosphatase were evaluated using the conventional synthetic substrates sodium beta glycerophosphate (betaGP) and p-nitrophenyl phosphate (PNPP) and physiologically occurring phosphate esters of carbohydrates, vitamins and nucleotides. The extent of hydrolysis varied depending on the substrates; phosphate esters of vitamins and carbohydrates were in general poor substrates. Kinetic analysis revealed the presence of two components of the enzyme for all the substrates. Component I had low Km and low Vmas. Opposite was true for component II. The Km values were generally high for betaGP, PNPP and adenosine diphosphate (ADP). Amongst the nucleotides substrates AMP showed high affinity i.e. low Km. The increase in enzyme activity in general at high substrate concentration seems to be due to substrate binding and positive cooperativity. AMP which showed highest affinity was inhibitory at high concentration beyond 1 mM. The results suggest that in situ the nucleotides may be the preferred substrates for acid phosphatase.  相似文献   

9.
1. The kinetics of mitochondrial mammalian pyruvate dehydrogenase multienzyme complex (PDHC) is studied by the formation of CO2 using tracer amounts of [1-14C]pyruvate. It is found that the Hill plot results in a (pseudo-)cooperativity with a transition of n-1----3 at a pyruvate concentration about Ks. 2. Addition of L-carnitine, octanoate, palmitoyl-CoA or palmitate + L-carnitine + fatty acid-binding protein results in a Hill coefficient of n = 2 following the kinetics of pyruvate oxidation. 3. Addition of fatty acid-binding protein to an assay system oxidizing palmitate in presence of L-carnitine alters the pattern of the kinetics in the Hill plot so that an apparently lower level of L-carnitine is necessary for the reaction course of beta-degradation. 4. It is concluded that beta-degradation is a coordinated, multienzyme-complex based mechanism tightly linked to citric acid cycle and it is proposed that L-carnitine is actively involved into the reaction and not only functioning as carrier-molecule for transmembrane transport.  相似文献   

10.
This paper is devoted to analysis of parameters of catalytical activity of CD45, the major transmembrane proteintyrosine phosphatase (PTP-ase) of the lymphocytes, isolated from plasma membranes of thymocytes of control and 0.5 Gy irradiated rats. CD45 catalytic features were evaluated using 0.2 mM sodium vanadate as the inhibitor and paranitrophenylphosphate (1-8 mM) and phosphotyrosine (1-6 mM) as, respectively, nonspecific and specific substrates. With the former, irradiation was shown to cause a decrease in Vmax but an increase in affinity. With phosphotyrosine both Vmax and affinity decreased. These data suggest that the exposure to radiation causes an increase in non-specific enzyme activity with a decrease in the ability to dephosphorylate the specific substrate. A study of cooperativity parameters shows that cooperativity between two phosphatase domains increased after irradiation. An analysis of the inhibitor kinetics showed that radiation caused a change of competitive inhibition by mixed one.  相似文献   

11.
The major secreted isoenzyme of human prostatic acid phosphatase (PAcP) (EC 3.1.3.2), which catalyses p-nitrophenyl phosphate (PNPP) hydrolysis at acid pH values, was found to have phosphotyrosyl protein phosphatase activity since it dephosphorylated three different phosphotyrosine-containing protein substrates. Several lines of evidence are presented to show that the phosphotyrosyl phosphatase and PAcP are the same enzyme. A highly purified PAcP enzyme preparation which contains a single N-terminal peptide sequence was used to test for the phosphotyrosyl phosphatase activity. Both activities comigrated during gel filtration by high performance liquid chromatography. Phosphotyrosyl phosphatase activity and PNPP acid phosphatase activity exhibited similar sensitivities to different effectors. Both phosphatase activities showed the same thermal stability. Specific anti-PAcP antibody reacted to the same extent with both phosphatase activities. PNPP acid phosphatase activity was competitively inhibited by the phosphotyrosyl phosphatase substrate. To characterize further the phosphotyrosyl phosphatase activity, the Km values using different phosphoprotein substrates were determined. The apparent Km values for phosphorylated angiotensin II, anti-pp60src immunoglobulin G and casein were in the nM range for phosphotyrosine residues, which was about 50-fold lower than the Km for phosphoserine residues in casein.  相似文献   

12.
The X-ray crystal structure of human prostatic acid phosphatase (PAP) in complex with a phosphate ion has been determined at 2.4 A resolution. This structure offers a snapshot of the final intermediate in the catalytic mechanism and does not support the role of Asp 258 as a proton donor in catalysis. A total of eight hydrogen bonds serve to strongly bind the phosphate ion within the active site. Bound PEG molecules from the crystallization matrix have allowed the identification of a channel within the molecule that likely plays a role in molecular recognition and in macromolecular substrate selectivity. Additionally, the structure of PAP in complex with a phosphate derivative, alpha-benzylaminobenzylphosphonic acid, a potent inhibitor (IC(50) = 4 nM), has been determined to 2.9 A resolution. This structure gives new insight into the determinants of binding hydrophobic ligands within the active site and allows us to explain PAP's preference for aromatic substrates.  相似文献   

13.
Acid phosphatase-1 (orthophosphoric monoester phosphohydrolase, acid optimum, EC 3.1.3.2), the major phosphatase in adult Drosophila melanogaster, has been purified to apparent homogeneity. The final product is a glycoprotein homodimer with a subunit molecular weight of about 50,000, as measured by its electrophoretic mobility in denaturing conditions on polyacrylamide gels containing sodium dodecyl sulfate. It has a turnover number of 1720 1-naphthyl phosphate molecules hydrolyzed/s by each acid phosphatase-1 molecule at 37 degrees C, pH 5.0. An average fly contains about 5 ng of enzyme. Pure acid phosphatase-1 displays heterogeneity in isoelectric focusing, with a major band at pH 5.3. The enzyme hydrolyzes a wide variety of phosphate monoesters, including AMP, glucose 6-phosphate, ATP, choline phosphate, or phosphoproteins. The maximum reaction rates are different for all substrates, and some substrates appear to inhibit the reaction at high substrate concentrations. The Michaelis constants for 1-naphthyl phosphate and p-nitrophenyl phosphate are 79 microM and 68 microM, respectively, at pH 5.0 and 37 degrees C. The optimum pH level for 1-naphthyl phosphate is 4.5. Acid phosphatase-1 is inhibited by L(+)-tartrate (but not D(-)-tartrate), phosphate, and fluoride. The reaction rate increases 2.1-fold for every 10 degrees C rise in temperature. Above 48 degrees C, the rate of thermal denaturation is greater than the rate of the enzyme reaction.  相似文献   

14.
In addition to kallikrein hK3, a serine protease generally reported as PSA (prostate-specific antigen), at least two other enzymes in human seminal plasma also cleave synthetic peptidyl substrates derived from the sequence of human semenogelins. We have identified one of these as prostatic acid phosphatase (PAP), a major component of prostatic fluid whose physiological function is unclear. The other is a high Mr basic protein present at low concentrations in seminal plasma and that remains to be characterized. PAP was purified to homogeneity from freshly ejaculated seminal plasma. Its N-terminal sequence and its phosphatase properties (hydrolysis of para-nitrophenylphosphate at low pH) were determined, and its inhibition by sodium fluoride measured. Both purified and commercial PAP also had amidolytic activity on peptide substrates derived from the semenogelin sequence at neutral and slightly basic pH. The k(cat)/K(m) values were in the 10(2)-10(3) m(-1) x s(-1) range using fluorogenic semenogelin-derived substrates whose peptidyl moiety included cleavage sites that had been identified ex vivo. PAP cleavage sites differed from those of hK3 and were mainly at P1 = Gln residues or between residues bearing hydroxyl groups. PAP amidolytic activity was poorly inhibited by all currently used wide spectrum proteinase inhibitors. Only 3-4 dichloroisocoumarin and benzamidine inhibited purified PAP. Purified human semenogelin was cleaved by purified and commercial PAP at neutral pH; the two main cleavage sites were at Tyr292 and Ser170 (semenogelin I sequence), only the former has been identified ex vivo by analysis of seminal plasma.  相似文献   

15.
Hydrolysis of adenosine 5'-triphosphate (ATP) and p-nitrophenyl phosphate by the hydrogen ion-transporting potassium-stimulated adenosine triphosphatase (H,K-ATPase) was investigated. Hydrolysis of ATP was studied at pH 7.4 in vesicles treated with the ionophore nigericin. The kinetic analysis showed negative cooperativity with one high affinity (Km1 = 3 microM) and one low affinity (Km2 = 208 microM) site for ATP. The rate of hydrolysis decreased at 2000 microM ATP indicating a third site for ATP. When the pH was decreased to 6.5 the experimental results followed Michaelis-Menten enzyme kinetics with one low affinity site (Km = 116 microM). Higher concentrations than 750 microM ATP were inhibitory. Proton transport was measured as accumulation of acridine orange in vesicles equilibrated with 150 mM KCl. The transport at various concentrations of ATP in the pH interval from 6.0 to 8.0 correlated well with the Hill equation with a Hill coefficient between 1.5-1.9. The concentration of ATP resulting in half-maximal transport rate (S0.5) increased from 5 microM at pH 6.0 to 420 microM at pH 8.0. At acidic pH the rate of proton transport decreased at 1000 microM ATP. The K+-stimulated p-nitrophenylphosphatase (pNPPase) activity resulted in a Hill coefficient close to 2 indicating cooperative binding of substrate. The pNPPase was noncompetitively inhibited by ATP and ADP; half-maximal inhibition was obtained at 2 and 100 microM, respectively. Phospholipase C-treated vesicles lost 80% of the pNPPase activity, but the Hill coefficient did not change. These kinetic results are used for a further development of the reaction scheme of the H,K-ATPase.  相似文献   

16.
The interaction of human organ alkaline phosphatases (orthophosphoric-monoester phosphohydrolases (alkaline optimum), EC 3.1.3.1) with sugars was studied. Hexosamines, N-acetylneuraminic acid (NANA or sialic acid), N-acetylmuramic acid and N-acetylglycolylneuraminic acid inhibited human organ alkaline phosphatase activities. Of these, sialic acid was the most effective inhibitor. The pH profiles for the enzymes in the absence and presence of sialic acid were similar. The sialic acid - enzyme complex was more heat stable than the free enzyme between 20 and 45 degrees C. Lineweaver-Burk plots of 1/v versus 1/S at various concentrations of sialic acid showed intersecting straight lines indicating that the mechanism of inhibition was a mixed type. The Ki value obtained from the plots of 1/v versus the square of sialic acid concentration was 0.07 mM for the hepatic, sialidase-treated hepatic, and intestinal alkaline phosphatases. The respective Hill coefficients varied somewhat with the alkaline phosphatase isoenzyme. Hyperbolic curves were obtained when the percentage of remaining activity was plotted against the substrate concentration at different concentrations of sialic acid. The Hill coefficient was lowered in the presence of sialic acid. The sialidase-treated hepatic enzymes used gave the most effective conversion. Partial denaturation of the enzyme with urea, or pronase digestion had a little if any effect on the sialic acid inhibition with constant time.  相似文献   

17.
Cooperativity in ATP hydrolysis by GroEL is increased by GroES.   总被引:3,自引:0,他引:3  
T E Gray  A R Fersht 《FEBS letters》1991,292(1-2):254-258
The kinetics of ATP hydrolysis by the 'molecular chaperone' GroEL and the inhibition of this hydrolysis by GroES have been studied in more detail. It is shown that the hydrolysis of ATP by GroEL is cooperative with respect to ATP with a Hill coefficient of 1.86 (+/- 0.13). In the presence of GroES, there is an increase in the degree of cooperativity with a Hill coefficient of 3.01 (+/- 0.18). The observed cooperativity is not due to dissociation of the GroEL oligomer into smaller units but more probably involves structural changes within the GroEL oligomer.  相似文献   

18.
Tyapochkin E  Cook PF  Chen G 《Biochemistry》2008,47(45):11894-11899
Cytosolic sulfotransferase (SULT)-catalyzed sulfation regulates biosignaling molecular biological activities and detoxifies hydroxyl-containing xenobiotics. The universal sulfuryl group donor for SULTcatalyzed sulfation is adenosine 3'-phosphate 5'-phosphosulfate (PAPS). The reaction products are a sulfated product and adenosine 3',5'-diphosphate (PAP). Although the kinetics has been reported since the 1980s,SULT-catalyzed reaction mechanisms remain unclear. Human SULT1A1 catalyzes the sulfation of xenobiotic phenols and has very broad substrate specificity. It has been recognized as one of the most important phase II drug-metabolizing enzymes. Understanding the kinetic mechanism of this isoform is important in understanding drug metabolism and xenobiotic detoxification. In this report, we investigated the SULT1A1-catalyzed phenol sulfation mechanism. The SULT1A1-catalyzed reaction was brought to equilibrium by varying substrate (1-naphthol) and PAPS initial concentrations. Equilibrium constants were determined. Two isotopic exchanges at equilibrium ([14C]1-naphthol <=>[14C]1-naphthyl sulfate and[35S]PAPS<=>[35S]1-naphthyl sulfate) were conducted. First-order kinetics, observed for all the is otopic exchange reactions studied over the entire time scale that was monitored, indicates that the system was truly at equilibrium prior to addition of an isotopic pulse. Complete suppression of the 35S isotopic exchange rate was observed with an increase in the levels of 1-naphthol and 1-naphthyl sulfate in a constant ratio,while no suppression of the 14C exchange rate was observed with an increase in the levels of PAPS and PAP in a constant ratio. Data are consistent with a steady state ordered kinetic mechanism with PAPS and PAP binding to the free enzyme.  相似文献   

19.
In the present work we have partially characterized an ecto-phosphatase activity in Crithidia deanei, using viable parasites. This enzyme hydrolyzed p-nitrophenylphosphate at a rate of 3.55 +/- 0.47 nmol Pi/h x 10(8) cells. The dependence on p-NPP concentration shows a normal Michaelis-Menten kinetics for this phosphatase activity and the value of the apparent Km for p-NPP was 5.35 +/- 0.89 mM. This phosphatase activity was inhibited by the product of the reaction, the inorganic phosphate. Experiments using classical inhibitors of acid phosphatases, such as ZnCl2 and sodium fluoride, as well as inhibitors of phosphotyrosine phosphatase, such as sodium orthovanadate and ammonium molybdate, showed a decrease in this phosphatase activity, with different patterns of inhibition.  相似文献   

20.
Summary Acid phosphatases of the rat ventral prostate were studied cytochemically using different substrates. The results were compared to findings on isoelectric focussing gels stained for acid phosphatase activity. This is a highly specific and reproducible method which allows the distinction between secretory androgen-dependent and lysosomal acid phosphatases. Activity of lysosomal acid phosphatase was increased after castration, while the activity of the secretory enzyme gradually decreased after androgen deprivation. None of the substrates tested was selectively hydrolyzed by either secretory or lysosomal acid phosphatase. Phenylphosphate, creatine phosphate and choline phosphate were found to be inappropriate substrates for histochemical purposes, however, reproducible results were obtained with -naphthylphosphate, -glycerophosphate and p-nitrophenylphosphate. The method of isoelectric focussing (pH range 4.0–8.0) of enzymes with subsequent histochemical staining demonstrated lysosomal enzymes at pH 7.9 and 8.2 respectively. Small amounts of identical enzymes were found in liver, kidney, blood or epididymis. Secretory acid phosphatases were focussed at pH 5.5, 5.6, 5.65 and 7.15. Similar enzymes have been identified in epididymis, kidney, liver and pancreas. These results indicate that 1) at present no specific substrate for prostatic secretory or lysosomal acid phosphatases is available and 2) that no prostate-specific prostatic acid phosphatase (PAP) exists in the rat ventral prostate.Supported by the Deutsche Forschungsgemeinschaft (Au 48/6)  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