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 structural basis of anomalous kinetics of rabbit liver aryl sulfatase A   总被引:1,自引:0,他引:1  
Rabbit liver aryl sulfatase A (aryl sulfate sulfohydrolase, EC 3.1.6.1) is inactivated during the hydrolysis of nitrocatechol sulfate and the rate of formation of turnover-modified aryl sulfatase A depends on the initial velocity of the enzymatic reaction. Organic solvents such as ethanol and dioxane favor the anomalous kinetic behavior. The turnover-modified enzyme can apparently be reactivated by arsenate, phosphate, pyrophosphate, and sulfate in the presence of nitrocatechol sulfate. The apparent dissociation constants of these ions in the reactivation of the enzyme are similar to their Ki values. Sulfite, which is a competitive inhibitor, does not reactivate the turnover-modified enzyme. Thus, all known activators are competitive inhibitors but not all competitive inhibitors are effective as activators. Inactivation of aryl sulfatase A during hydrolysis of 35S-labeled substrate at pH values near the pH optimum (pH 5–6) is accompanied by the incorporation of radioactivity into the protein molecule and the turnover-modified enzyme is thereby covalently labeled. The stoichiometry of the incorporation of radioactivity corresponds to 2 g atom of sulfur per mole of enzyme monomer, or 1 g atom of sulfur per equivalent peptide chain. It is also shown that isolated turnover-modified rabbit liver aryl sulfatase A has lost approximately 76% of its secondary structure as compared to the native enzyme. The specific activity of the inactive enzyme is also decreased by 82%. Turnover-modified rabbit liver aryl sulfatase A is partially reactivated by sulfate ions in the presence of nitrocatechol sulfate. However, circular dichroism measurements and fluorescence spectra of the isolated “reactivated” turnover-modified enzyme indicate only a further loss of secondary structure. The specific activity of this “reactivated” enzyme is in fact decreased. The loss in secondary structure and the enzyme activity of the “reactivated” aryl sulfatase A is prevented in the presence of sulfate ions. Turnover-modified rabbit liver aryl sulfatase A behaves as a very fragile molecule.  相似文献   

2.
In a recent communication, A. Waheed and R. L. Van Etten (1979, Arch. Biochem. Biophys. 195, 248) showed that the sulfatase A of rabbit liver (arylsulfate sulfohydrolase, EC 3.1.6.1), which becomes inactivated as it catalyzes the hydrolysis of substrate, covalently incorporates 35S from nitrocatechol [35S]sulfate during this reaction and at the same time loses most of its secondary structure in solution. Circular dichroism spectra presented here for the native and turnover-modified forms of the sulfatase A of ox liver indicate no difference in the region of the spectrum below 240 nm associated with polypeptide backbone contributions or in the region from 350-250 nm associated with the side-chain chromophore transitions. In addition no differences were evident for the two forms of the ox liver enzyme from ultraviolet absorbance and fluorescence spectroscopy measurements. From these data we conclude that, in contrast to the situation with the rabbit enzyme, there is no loss of secondary structure associated with inactivation of ox liver sulfatase A in the course of enzymic catalysis.  相似文献   

3.
Aryl sulfatase A (aryl sulfate sulfohydrolase EC 3.1.6.1) has been purified > 10,000-fold from rabbit liver; by disc gel electrophoresis the enzyme appears homogeneous. Various properties of the enzyme have been determined and comparisons are made with other aryl sulfatases. Sodium dodecyl sulfate gel electrophoresis indicates that the enzyme is made up of monomers of molecular weight ~ 70,000. At pH 7.4 the enzyme exists as a dimer whereas a tetrameric form predominates at pH 4.8.The enzyme exhibits the anomalous kinetics often observed with aryl sulfatase A from mammalian tissues (the enzyme is modified to an inactive form while degrading substrate and the inactive form can be reactivated by sulfate ion). The enzyme activity has been studied under a variety of reaction conditions. Two pH optima are observed and neither enzyme concentration or changes in ionic strength appear to have an effect on the relative magnitudes of the optima. Aryl sulfatase A is competitively inhibited by potassium sulfate, potassium phosphate, and sodium sulfite (Ki = 2.9 × 10?3 M, 3.4 × 10?5 M, and 1.1 × 10?6 M, respectively). Kinetic constants for some substituted phenyl sulfate esters have been determined. The variation in V is not consistent with a reaction mechanism involving a rate-limiting breakdown of a common intermediate.The inactive (modified) form of the enzyme has been isolated from reaction mixtures containing aryl sulfatase A and substrate. A procedure is presented for determining the relative amount of modified and native enzyme in these preparations. In the presence of substrate, sulfate displaces the equilibrium between native and modified enzyme in favor of native enzyme. In the absence of substrate neither sulfate or phosphate have an effect on the equilibrium. A study is made of the temperature dependence of the process in which the modified enzyme is converted back to native enzyme. The relatively small entropy of activation for the conversion of the modified to the native form (ΔS3 = ?8 cal/mole deg) does not seem to be consistent with a major modification of protein conformation.  相似文献   

4.
The polymerization of aryl sulfatase A (aryl sulfate sulfohydrolase, EC 3.1.6.1) has been studied by frontal gel chromatography on Sephadex G-200 and Bio-Gel A-5m under various conditions of pH, ionic strength, and temperature. The aryl sulfatase A molecule exists as a monomer and as a dimer at pH 7.5 and pH 4.5, respectively. The extent of dissociation is markedly pH-, protein concentration-, and ionic strength-dependent. Only a small effect of temperature was observed. The enthalpy change (ΔHo) for the dissociation was ?2.5 ± 1 kcal/mol at pH 5.5–5.6, and the entropy change for dissociation of the enzyme dimer to two monomeric units was ?47 cal mol?1 deg?1. Sulfate ion has little effect on the extent of dissociation of the enzyme at pH 5.6. The present studies suggest that the dissociation of rabbit liver aryl sulfatase A is regulated by the ionization of amino acid residues whose apparent pK is between pH 5 and 6. The driving force for the association of the subunits of the enzyme is primarily ionic and/or ionic/hydrogen bond formation. The small enthalpy change and the fact that dissociation is strongly favored by an increase in the ionic strength suggest that hydrophobic interactions play only a minor role in stabilizing the dimeric quaternary structure relative to the monomeric state. The monomeric form of the enzyme exhibits the anomalous kinetics often observed with sulfatase A but the dimer does not show anomalous kinetics. Since aryl sulfatase A is probably in the dimeric form in the lysosome, the anomalous kinetics of the enzyme are unlikely to be of physiological importance in the intact lysosome.  相似文献   

5.
Rabbit liver aryl sulfatase A (aryl-sulfate sulfohydrolase, EC 3.1.6.1) is a glycoprotein containing 4.6% carbohydrate in the form of 25 residues of mannose, seven residues of N-acetylglucosamine, and three residues of sialic acid per enzyme monomer of molecular weight 140 000. Each monomer consists of two equivalent polypeptide chains. The protein has a relatively high content of proline, glycine and leucine, and the amino acid composition of rabbit liver aryl sulfatase A is similar to that of other known liver sulfatases. Rabbit liver aryl sulfatase A catalyzes the hydrolysis of a wide variety of sulfate esters, although it appears possible that cerebroside sulfate is a physiological substrate for the enzyme because the Km is very low (0.06 mM). The turnover rate for hydrolysis of nitrocatechol sulfate or related synthetic substrates is much higher than the rate with most naturally occurring sulfate esters such as cereroside sulfate, steroid sulfates, L-tyrosine sulfate or glucose 6-sulfate. However, the turnover rate with ascorbate 2-sulfate is comparable to the rates measured using most synthetic substrates. These results are discussed in relationship to several previously described sulfatase enzymes which were claimed to have unique specificities.  相似文献   

6.
A detailed study of the pH dependence of the Michaelis-Menten constants (V and Km) of aryl sulfatase A (EC 3.1.6.1) from rabbit liver indicates that at least two functional groups (pK's ~4.3 and ~7 in the enzyme-substrate complex) participate in the enzymic degradation of substrate. Aryl sulfatase A is inactivated by diethyl pyrocarbonate (ethoxyformic anhydride). The enzyme that has been modified with this reagent can in turn be reactivated by treatment with hydroxylamine. The pH dependence of inactivation reveals a reactive group having a pK of 6.5–7.0. The results indicate that at least one histidine plays an important catalytic role in rabbit liver aryl sulfatase A, consistent with the results of earlier workers who employed diazotized sulfanilic acid. Phosphate ion, a competitive inhibitor, partially protects the enzyme from inactivation by diethyl pyrocarbonate whereas sulfate ion, also a competitive inhibitor, increases the rate of inactivation by diethyl pyrocarbonate. This result is of particular significance in view of the anomalous kinetics of aryl sulfatase A. The kinetic effects of even small amounts of sulfate ion impurities in many commercial sulfate ester substrate preparations is also discussed.  相似文献   

7.
The monomeric form of rabbit liver aryl sulfatase A (aryl sulfate sulfohydrolase, EC 3.1.6.1) was covalently coupled to CNBr-activated Sepharose and the catalytic properties of the covalently coupled monomer subunit were examined. The immobilized subunit showed one pH optimum near pH 5.6 which appears to be the characteristic pH optimum of the monomer. The enzyme-Sepharose complex exhibited the characteristic anomalous kinetic behavior at pH 5.5 but there was no turnover-induced inactivation of the immobilized enzyme at pH 4.5. The covalently coupled subunit column was examined for its ability to act as a subunit affinity chromatography medium. It was found that dissolved aryl sulfatase A was removed from solution at pH 4.5 and pH 5.0, I = 0.2, and became associated with the affinity column of Sepharose-aryl sulfatase A. The retained subunit of the enzyme could subsequently be quantitatively eluted with 0.2 m Tris-HCl, pH 7.5. Extraneous protein such as bovine serum albumin did not measureably affect the rate or equilibrium for association of the enzyme to the covalently bound subunit. The extent of binding of the enzyme to the affinity column was found to be strongly dependent on the time of equilibration and on the pH. About 90% of the enzyme was retained after 24 h at pH 5.0, I = 0.2. Under otherwise comparable conditions, use of Sepharose-6MB resulted in slightly faster association than did Sepharose-4B. Under the experimental conditions employed, the total capacity of the affinity column was approx 50% of the total aryl sulfatase A coupled to the Sepharose. The rabbit liver subunit column also permits the purification of several other aryl sulfatase A enzymes. Thus, the subunit affinity column provides a simple, convenient, and rapid procedure for the isolation of most mammalian aryl sulfatase A enzymes as well as for studying inter- and intraspecific subunit association interactions.  相似文献   

8.
We have reported that the monovalent ionophore monensin causes undersulfated chondroitin sulfate biosynthesis in cultured chondrocytes. In order to clarify the mechanism of this diminished sulfation, we have measured the rate of incorporation of sulfate into chondrocytes and assayed the cellular ATP levels. We have also measured sulfatase activity, the incorporation of 35SO4 into 3′-phosphoadenosine 5′-phospho[35S]sulfate and endogenous sulfotransferase activity in the cell-free extracts. We find that: (1) The incorporation of 35SO4 into the free sulfate pool in chondrocytes was not inhibited by monensin. (2) The ATP levels of monensin-treated chondrocytes were the same as control cells. (3) There was no sulfatase activity in both control and monensin-treated chondrocytes. (4) Enzymatic analyses revealed that 35SO4 incorporation into 3′-phosphoadenosine 5′-phospho[35S]sulfate and subsequent sulfotransferase activity were not inhibited in the presence of monensin. At present the most tenable hypothesis to account for monensin causing undersulfated chondroitin sulfate synthesis is that the ionophore impairs the access of proteoglycans to the sulfotransferases in the luminal walls of the Golgi structures.  相似文献   

9.
The activation of the preemergent herbicide 2-(2,4-dichlorophenoxy)ethyl sulfate (Crag herbicide) is initiated by soil microorganisms that are presumed to act by removing the ester sulfate group via some type of sulfatase enzyme. An enrichment technique with the herbicide as the sole source of sulfur led to the isolation of several pure cultures that could produce 2-(2,4-dichlorophenoxy)ethanol from the herbicide. One of these, a strain of Pseudomonas putida, was particularly active. Polyacrylamide gel zymograms of extracts of cells grown on nutrient broth showed the presence of three secondary and three primary alkylsulfatases. One of the latter enzymes was active toward Crag herbicide as well as sodium dodecyl sulfate. Maximum activity was obtained in the late-stationary phase of growth, and enzyme yields were not affected by either the presence or the absence of the herbicide in the growth medium. The enzyme was purified 2,670-fold to homogeneity by a combination of streptomycin sulfate treatment, heat treatment, and column chromatography on DEAE-cellulose, Sephacryl 200-S, and butyl agarose. The pure enzyme was tetrameric (molecular weight, 295,000) and most active at pH 6.0. Saturation kinetics with inhibition by excess substrate were observed for Crag herbicide and octyl sulfate. 2-Butox-yethyl sulfate was a relatively poor substrate, and dodecyltriethoxy sulfate was not hydrolyzed at all. Enzymatic hydrolysis of each substrate in the presence of H218O led to incorporation of 18O exclusively into SO42− ions in all three cases. The Crag herbicide sulfatase therefore acts by cleaving the O-S bond of the C-O-S ester linkage, in contrast with other alkylsulfatases acting on long-chain alkyl sulfates.  相似文献   

10.
Abstract– The enzymatic hydrolysis by brain homogenate of the sulfate esters of estrone, pregnenolone, dehydroepiandrosterone, testosterone, cholesterol and p-nitrophenol was studied. With homogenate of young rat brain, the pH optima of estrone sulfatase 4 4 The term steroid sulfatase is used as a general name for the enzyme(s) which hydrolyzes the sulfate ester of a steroid. Simplified terms, such as estrone sulfatase, instead of the more formal terms, such as estrone sulfate sulfohydrolase, have been used throughout.
and arysulfatase C (p-nitrophenyl sulfate as substrate) were 8.2 and all other steroid sulfatases had pH optima at 6.6. Apparent Kms for these steroid sulfates were widely different. The highest Km value was 32.2 μm for estrone sulfate and the lowest was 0.66 μm for testosterone sulfate; the Km for p-nitrophenyl sulfate was 30 fold higher than for estrone sulfate. Specific activity was also highest with estrone sulfatase and lowest with testosterone sulfatase; specific activity with aryl sulfatase C was over 3 fold higher than with estrone sulfatase. Estrone sulfatase activity was inhibited noncompetitively by sulfate esters of dehydroepiandrosterone, pregnenolone, and cholesterol; on the other hand, other steroid sulfatases were inhibited by these latter three sulfates competitively. Developmental changes of these sulfohydrolase activities in rat brain were almost identical with the exception of testosterone sulfatase activity; the latter sulfatase had a peak activity at 30 days old, while all other sulfatase had a peak at 20 days old. Thermal stability of all these activities was identical. Testosterone sulfatase activity in neurological mouse mutants, jimpy, msd, and quaking mice, was less than one half of littermate controls, while other steroid sulfatase levels in these mutants' brain were normal. All sulfatase activities were diminished in the brain of a metachromatic leukodystrophy patient with multiple sulfatase deficiency. The brains of classical metachromatic leukodystrophy patients contained normal levels of all steroid sulfatases and arylsulfatase C, with the single exception of testosterone sulfatase which level was less than 50% of control.  相似文献   

11.
An aryl sulfatase of unusual specificity has been isolated from the liver of marine mollusk Littorina kurila. It hydrolyzes p-nitrophenyl sulfate, does not affect the natural fucoidan, and catalyzes splitting off the sulfate group in position C4 of xylose residues within the carbohydrate chains of holostane triterpene glycosides from sea cucumbers. The properties of the enzyme were studied at pH 5.4. The protein is homogeneous according to electrophoresis and has M 45 ± 1 kDa. The semiinactivation time of the enzyme at 60°C is 20 min, and its K m value for the hydrolysis of p-nitrophenyl sulfate is 8.7 ± 1 mM. It was shown that natural sulfated polyhydroxysteroids inhibit activity of the sulfatase; their I 50 values depend on their structures and are within the range from 10?3 to 10?5 M.  相似文献   

12.
Primary Alcohol Sulfatase in a Pseudomonas Species   总被引:9,自引:4,他引:5       下载免费PDF全文
An ammonium sulfate-precipitated fraction from cell-free extracts of Pseudomonas C12B grown on a medium containing sodium dodecyl sulfate (SDS) contained alkyl sulfatase increased fourfold in specific activity over the crude. Optimal pH (7.5) and temperature (70 C) for sulfate release were determined with SDS labeled with radioactive sulfur (SDS(35)) as test substrate. Phosphate, arsenate, and certain heavy metal ions inhibited desulfation, whereas Mg(++) and Mn(++) stimulated activity of preparations which had been dialyzed against ethylenediaminetetraacetic acid. Dodecanol was recovered in semiquantitative yield from reaction mixtures containing enzyme and SDS(35). Aryl sulfates, secondary alcohol sulfates, and a phenoxyethyl sulfate failed to serve as substrate for this enzyme.  相似文献   

13.
Steroid sulfatase of human placenta has been solubilized by treatment of the microsomal fraction with an amphoteric surface active agent, Miranol H2M and ultrasound. Criteria of solubility include non-sedimentation of the activity following centrifugation at 160,000 × g, its retention on Sepharose 6B and a single peak of activity after polyacrylamide gel electrophoresis. Enzyme activity was located in the same gel fractions for the two substrates tested; cholesterol sulfate and dehydroisoandrosterone sulfate. The addition of dithiothreitol was found necessary to maintain the stability of the enzyme indicating the presence of sulfhydryl groups in the molecule. A molecular weight of approximately 330,000 has been estimated from the elution volume of the enzyme system on a column of Sepharose 6B. It is believed that this protein represents a sulfatase enzyme complex composed of subunits with different specificities. From kinetic studies, a Km of 6.2 × 10?5M for the cleavage of dehydroisoandrosterone sulfate and a Km of 2 × 10?6M for the cleavage of cholesterol sulfate have been calculated.  相似文献   

14.
Often used to remove sulfate groups from carbohydrates, the regulatory properties of the aryl sulfatase from Helix pomatia remain little characterized. As many hydrolytic enzymes utilize exogenous metal ions in catalysis, the effect of various divalent metal ions on the sulfatase was investigated. Evidence for metal ion activation was collected, with Cd2+ being notable for effective activation. The enzyme was inhibited by Cu2+. The response of other common hydrolases to divalent metal ions was characterized. Activation by Cd2+ was not observed for chymotrypsin, rabbit liver esterase, or β-galactosidase. Instead, Cd was found to inhibit both the esterase and the galactosidase. Inhibition by Cu2+ and Zn2+ was also observed for some of these hydrolases.  相似文献   

15.
Carbonic anhydrases (CA) catalyze activated ester hydrolysis in addition to the hydration of CO2 to bicarbonate. They also show phosphatase activity with 4-nitrophenyl phosphate as substrate but not sulfatase with the corresponding sulfate. Here we prove that the enzyme is catalyzing the synthesis of cyclic diols from sulfate esters. 5-, 6- and 8-membered ring cyclic sulfates incorporating a neighboring secondary alcohol moiety were treated with CA II and yielded the corresponding cyclic diols. Inhibitory properties of obtained cyclic and original sulfate esters were then investigated on human carbonic anhydrase I (hCA I), hCA II, hCA IV and hCA VI (h?=?human isoform). KI-s of these compounds ranged between 32.7–423 μM against hCA I, 2.13–32.4 μM against hCA II, 13.7–234 μM against hCA IV and 76–278 μM against CA VI, respectively. The sulfatase activity of CA with such esters is amazing considering the fact that 4-nitrophenyl-sulfate is not a substrate of these enzymes.  相似文献   

16.
Regulation of the synthesis of several enzymes of sulfur metabolism in Neurospora is a function of both metabolic regulation and the genetic control exerted by the cys-3 and scon regulatory genes. Additional control mechanisms appear to regulate the synthesis of choline sulfatase and aryl sulfatase in different developmental stages of the life cycle. The metabolic regulation of enzyme synthesis in conidia differs from that which occurs in the mycelial stage. During conidial germination and mycelial outgrowth, the synthesis of these enzymes is not coordinate but begins at different times and occurs at different rates. A rapid and early synthesis of choline sulfatase was observed during conidial germination under derepressing conditions; furthermore, synthesis of the enzyme also occurred for a brief period in germinating conidia even in the presence of repressing levels of sulfate. The results of this study suggest that several enzymes of sulfur metabolism are independently controlled by a developmental system which is superimposed upon the cys-3 regulatory mechanism. It was also found that choline sulfatase undergoes rapid turnover while aryl sulfatase is a stable species.  相似文献   

17.
The synthesis of aryl sulfatase, choline-O-sulfate permease, and two distinct sulfate permeases are repressed by methionine, but the activity of these enzymes is not subject to feedback inhibition. The permease species, but not aryl sulfatase, are also regulated by dynamic turnover, displaying a functional half-life of approximately 2 hr. The rate of turnover of these permeases is not influenced by the presence of the end product, methionine. Development of sulfate permease activity occurs only by de novo synthesis which requires both a lifting of methionine repression and a functional cys-3 product. The turnover system for sulfate permease is not present in dormant conidia but appears to be synthesized relatively rapidly during germination. Preexisting conidial sulfate permease is lost by turnover during germination and outgrowth into the mycelial phase, during which both permease species are synthesized anew, although the high affinity system contributes most of the total activity in growing mycelia.  相似文献   

18.
Ensifer (Sinorhizobium) meliloti is a nitrogen-fixing α-proteobacterium able to biosynthesize the osmoprotectant glycine betaine from choline sulfate through a metabolic pathway that starts with the enzyme choline-O-sulfatase. This protein seems to be widely distributed in microorganisms and thought to play an important role in their sulfur metabolism. However, only crude extracts with choline sulfatase activity have been studied. In this work, Ensifer (Sinorhizobium) meliloti choline-O-sulfatase was obtained in a high degree of purity after expression in Escherichia coli. Gel filtration and dynamic light scattering experiments showed that the recombinant enzyme exists as a dimer in solution. Using calorimetry, its catalytic activity against its natural substrate, choline-O-sulfate, gave a kcat=2.7×10?1 s?1 and a KM=11.1 mM. For the synthetic substrates p-nitrophenyl sulfate and methylumbelliferyl sulfate, the kcat values were 3.5×10?2 s?1 and 4.3×10?2 s?1, with KM values of 75.8 and 11.8 mM respectively. The low catalytic activity of the recombinant sulfatase was due to the absence of the formylglycine post-translational modification in its active-site cysteine 54. Nevertheless, unmodified Ensifer (Sinorhizobium) meliloti choline-O-sulfatase is a multiple-turnover enzyme with remarkable catalytic efficiency.  相似文献   

19.
When cell-free preparations of Chlorella pyrenoidosa Chick (Emerson strain 3) form thiosulfate from labeled sulfate, another radioactive compound also appears. This compound has been isolated in quantity and is shown to be identical with adenosine-3′-phosphate-5′-phosphosulfate (PAPS) on the basis of its chromatographic and electrophoretic behavior, chemical composition, sensitivity to selective degradative enzymes, and its ability to serve as a substrate for rat liver aryl sulphotransferase. In addition, as expected for PAPS, the compound on mild acid treatment yields all of its radioactive sulfur as sulfate, and is converted to a compound identical with adenosine-3′,5′-diphosphate (PAP). Replacement of sulfate and ATP by this PAP35S in the usual incubation mixture yields the same product, thiosulfate, which can be isolated as such or detected as acid-volatile radioactivity. This conversion of PAP35S to thiosulfate still requires the addition of Mg2+ and a reductant such as 2,3-dimercaptopropan-1-ol (BAL). The cause of our previous result that high concentrations of ATP inhibit thiosulfate formation from sulfate can be ascribed to a small amount of PAP contaminating the ATP preparations, since PAP proves to be an exceedingly effective inhibitor of the conversion of PAP35S to thiosulfate. Sulfate reduction to thiosulfate by Chlorella extracts is discussed and compared with similar systems from other organisms.  相似文献   

20.
Mitochondria that have been purified from cells of light-grown wild-type Euglena gracilis Klebs var. bacillaris Cori or dark-grown mutant W10BSmL and incubated with 35SO4(2-) and ATP accumulate a labeled compound in the surrounding medium. This compound is also labeled when mitochondria are incubated with [14C]tyrosine and nonradioactive sulfate under the same conditions. This compound shows exact coelectrophoresis with synthetic tyrosine O-sulfate at pH 2.0, 5.8, and 8.0, and yields sulfate and tyrosine on acid hydrolysis. Treatment with aryl sulfatase from Aerobacter aerogenes yields sulfate and tyrosine but no tyrosine methyl ester; no hydrolysis of tyrosine methyl ester to tyrosine is observed under identical conditions, ruling out methyl esterase activity in the aryl sulfatase preparation. Thus the compound is identified as tyrosine O-sulfate. No tyrosine O-sulfate is found outside purified developing chloroplasts of Euglena incubated with 35SO4(2-) and ATP, but both chloroplasts and mitochondria accumulate labeled tyrosine-O-sulfate externally when incubated with adenosine 3'-phosphate 5'-phospho[35S]-sulfate (PAP35S). Since tyrosine does not need to be added, it must be provided from endogenous sources. Labeled tyrosine O-sulfate is found in the free pools of light-grown Euglena cells grown on 35SO4(2-) or in dark-grown cells incubated with 35SO4(2-) in light, but none is found in the medium after cell growth. No labeled tyrosine O-sulfate is found in Euglena proteins (including those in extracellular mucus) after growth or incubation of cells with 35SO4(2-) or after incubation of organelles with 35SO4(2-) and ATP or PAP35S, ruling out sulfation of the tyrosine in protein or incorporation of free-pool tyrosine O-sulfate into protein. The system forming tyrosine O-sulfate is membrane-bound and may be involved in transporting tyrosine out of the organelles.  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