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.
The electron spin resonance signals of wild type Chlamydomonas reinhardi and three mutant strains having impaired photosynthesis have been investigated. The wild type strain generates two different electron spin resonance signals. Signal I is obtained without illumination (i.e., dark signal) whereas signal II is generated preferentially only by red light. Signal I is missing from wild type cells that have been cultured in the dark, but it returns after these dark-grown cells have been illuminated. Chloroplast fragments obtained from the three mutant strains cannot photoreduce TPN. Two of the strains lack the dark signal I while the third strain has both signal I and signal II. Other studies have revealed that the two mutant strains which lack signal I give no Hill reaction but that they can photoreduce TPN if supplied with an artificial reductant. The mutant strain which has both electron spin resonance signals can carry out the Hill reaction, yet it too will not photoreduce TPN unless reductant is supplied. The electron spin resonance signals generated by the wild type and mutant strains are discussed in terms of the pathway of TPN photoreduction, and it is suggested that signal I is associated with one of the two light-dependent phases of this pathway.  相似文献   

2.
Folate-binding protein(s) from chronic myelogenous leukemia cells have been purified using acid dialysis, ammonium sulfate fractionation and affinity chromatography. The purified preparation which migrates as a single band on disc electrophoresis could be separated by DEAE agarose chromatography into two folate-binding proteins (binders I and II) which bind molar equivalents of folic acid. One binder (I) eluted from DEAE at 1 mM sodium phosphate, pH 6.0, and the other (II) at 100 mM sodium phosphate, pH 7.4. Analysis of the purified mixture, which contained more than 90% binder II, by sedimentation equilibrium centrifugation indicated a homogeneous protein with a calculated molecular weight of 44000. Antiserum raised against the purified mixture gave a single precipitin line by immunodiffusion against a preparation of partially purified cell lysate.Hydrolysis of the more acidic binder (II) with neuraminidase converted it to a weakly acidic protein similar to binder I suggesting that these binders are glycoproteins which differ in sialic acid content. With isoelectric focusing, the binding of folic acid would be demonstrated at pH 6.7, 7.3, 7.8 and 8.2 for binder I, and at pH 5.1, 5.8 and 6.5 for binder II. Binders I and II had equally high affinity for folic acid and dihydroflate, lower affinity of N5-methyl-tetrahydrofolate, and no apparent affinity for N5-formytetrahydrofolate or methotrexate.  相似文献   

3.
The phycobilin pigments were freshly extracted from Porphyra naiadum in the cold. At least two types of phycoerythrin (I and II) can be distinguished by electrophoresis, chromatography, and spectral characteristics. At pH 5.0 phycoerythrin II has a relatively large negative charge, while phycoerythrin I is nearly iso-electric. At pH 7.0, however, phycoerythrin I has the larger negative charge. Mobilities have been calculated by visual measurement of electrophoresis. Phycoerythrin II can be converted to phycoerythrin I by storing at pH 7.0. Chromatography indicates at least two types of phycocyanin as well.  相似文献   

4.
《Experimental mycology》1990,14(2):113-123
Trichospores of the fungus,Smittium culisetae (Trichomycetes), require a two-phase sequence of stimuli for sporangiospore extrusion: Phase I (20 mM KCl, pH 10) and Phase II (pH 7). Phases I and II correspond to conditions within the respective midgut and hindgut regions of mosquito larval hosts. In this study, trichospore pressure increased 1.0 MPa during Phase I but sporangiospores were not extruded. Potassium uptake was insufficient to account for the pressure increase. Sporangiospore extrusion from trichospores was stimulated by a drop in pH at Phase II. Lowering the osmotic potential of the Phase II solution with sucrose inhibited sporangiospore extrusion to varying degrees, with total inhibition occurring at osmotic potentials ≤ −2.7 MPa. At an osmotic potential of −2.1 MPa, the interwall layer at the trichospore apex swelled at the point of sporangial rupture without emergence of the sporangiospore. Interwall layer swelling during Phase II may rupture the sporangial wall of the trichospore and allow the sporangiospore to be extruded by its Phase I pressure.  相似文献   

5.
The binding of Cd(II) and Zn(II) to human serum albumin (HSA) and dog serum albumin (DSA) has been studied by equilibrium dialysis and 113Cd(II)-NMR techniques at physiological pH. Scatchard analysis of the equilibrium dialysis data indicate the presence of at least two classes of binding sites for Cd(II) and Zn(II). On analysis of the high-affinity class of sites, HSA is shown to bind 2.08 +/- 0.09 (log K = 5.3 +/- 0.6) and 1.07 +/- 0.12 (log K = 6.4 +/- 0.8) moles of Cd(II) and Zn(II) per mole of protein, respectively. DSA bound 2.02 +/- 0.19 (log K = 5.1 +/- 0.8), and 1.06 +/- 0.15 (log K = 6.0 +/- 0.2) moles of Cd(II) and Zn(II) per mole of protein, respectively. Competition studies indicate the presence of one high-affinity Cd(II) site on both HSA and DSA that is not affected by Zn(II) or Cu(II), and one high-affinity Zn(II) site on both HSA and DSA that is not affected by Cd(II) or Cu(II). 113Cadmium-HSA spectra display three resonances corresponding to three different sites of complexation. In site I, Cd(II) is most probably coordinated to two or three histidyl residues, site II to one histidyl residue and three oxygen ligands (carboxylate), while for the most upfield site III, four oxygens are likely to be involved in the binding of the metal ion. The 113Cd(II)-DSA spectra display only two resonances corresponding to two different sites of complexation. The environment around Cd(II) at sites I and II on DSA is similar to sites I and II, respectively, on HSA. No additional resonances are observed in any of these experiments and in particular in the low field region where sulfur coordination occurs. Overall, our results are consistent with the proposal that the physiologically important high-affinity Zn(II) and Cd(II) binding sites of albumins are located not at the Cu(II)-specific NH2-terminal site, but at internal sites, involving mostly nitrogen and oxygen ligands and no sulphur ligand.  相似文献   

6.
Carnosine complexes with copper(II) ions were studied with magnetic resonance techniques over a wide range of ligand to metal ratios at various pH values. Water proton relaxation rates increased with decreasing carnosine to copper ratios until a molar ratio of 48 was reached. Over the ratio range of 48–2 carnosine molecules per copper ion, the relaxation rate decreased so that in the 2:1 carnosine-copper(II) complex, the water-copper(II) distance was estimated to be 1.92 Å. Proton NMR studies revealed the broadening of imidazole proton lines at high mole ratios followed by other histidyl protons as the ratio decreased. The β-alanyl methylene protons were the last to be broadened by the addition of copper(II) ions. Carbon to copper(II) distances were determined for the carnosine to copper mole ratios of 500:1 and 5000:1. EPR spectra obtained at 93°K revealed the probable existence of four carnosine imidazoles as the sole coordinated ligands to copper(II) at high dipeptide-to-metal ratios (>10). At mole ratios below four, nuclear hyperfine lines characteristic of both monomeric and dimeric carnosine-copper(II) forms were observed. These results reveal that imidazole from carnosine is the sole ligand contributed to copper(II) for coordination over the pH range 5 to 7 at high carnosine to copper(II) ratios  相似文献   

7.
Mammalian tapasin (TPN) is a key member of the major histocompatibility complex (MHC) class I antigen presentation pathway, being part of the multi-protein complex called the peptide loading complex (PLC). Several studies describe its important roles in stabilizing empty MHC class I complexes, facilitating peptide loading and editing the repertoire of bound peptides, with impact on CD8+ T cell immune responses. In this work, the gene and cDNA of the sea bass (Dicentrarchus labrax) glycoprotein TPN have been isolated and characterized. The coding sequence has a 1329 bp ORF encoding a 442-residue precursor protein with a predicted 24-amino acid leader peptide, generating a 418-amino acid mature form that retains a conserved N-glycosylation site, three conserved mammalian tapasin motifs, two Ig superfamily domains, a transmembrane domain and an ER-retention di-lysine motif at the C-terminus, suggestive of a function similar to mammalian tapasins. Similar to the human counterpart, the sea bass TPN gene comprises 8 exons, some of which correspond to separate functional domains of the protein. A three-dimensional homology model of sea bass tapasin was calculated and is consistent with the structural features described for the human molecule. Together, these results support the concept that the basic structure of TPN has been maintained through evolution. Moreover, the present data provides information that will allow further studies on cell-mediated immunity and class I antigen presentation pathway in particular, in this important fish species.  相似文献   

8.
Brauer D  Teel MR 《Plant physiology》1982,70(3):723-727
Kinetics of two molecular forms of K-dependent citrate dehydrase in maize (Zea mays L.) are reported. The isozymes, designated CD I and CD II, were found to be compartmented in mitochondria and cytosol, respectively.

CD I exhibited hyperbolic kinetics with respect to both citrate and potassium with Km 2.3 and 12 millimolar, respectively. Maximum velocity was 0.38 micromole of trans-aconitic acid per minute per milligram protein. The pH optimum was 7.2. trans-aconitic synthesis by CD I is regulated by both citrate concentration and pH.

CD II exhibited hyperbolic kinetics with respect to citrate (Km 0.6 millimolar) but sigmoidal kinetics with respect to potassium. trans-aconitic acid synthesis by CD II is regulated by potassium. This may account for the positive correlation between leaf potassium and trans-aconitic acid in certain grasses (Clark 1968 Crop Sci 8: 165).

  相似文献   

9.
Carnosine complexes with copper(II) ions were studied with magnetic resonance techniques over a wide range of ligand to metal ratios at various pH values. Water proton relaxation rates increased with decreasing carnosine to copper ratios until a molar ratio of 48 was reached. Over the ratio range of 48–2 carnosine molecules per copper ion, the relaxation rate decreased so that in the 2:1 carnosine-copper(II) complex, the water-copper(II) distance was estimated to be 1.92 Å. Proton NMR studies revealed the broadening of imidazole proton lines at high mole ratios followed by other histidyl protons as the ratio decreased. The β-alanyl methylene protons were the last to be broadened by the addition of copper(II) ions. Carbon to copper(II) distances were determined for the carnosine to copper mole ratios of 500:1 and 5000:1. EPR spectra obtained at 93°K revealed the probable existence of four carnosine imidazoles as the sole coordinated ligands to copper(II) at high dipeptide-to-metal ratios (>10). At mole ratios below four, nuclear hyperfine lines characteristic of both monomeric and dimeric carnosine-copper(II) forms were observed. These results reveal that imidazole from carnosine is the sole ligand contributed to copper(II) for coordination over the pH range 5 to 7 at high carnosine to copper(II) ratios  相似文献   

10.
Lakes formed because of coal mining are characterized by low pH and high concentrations of Fe(II) and sulfate. The anoxic sediment is often separated into an upper acidic zone (pH 3; zone I) with large amounts of reactive iron and a deeper slightly acidic zone (pH 5.5; zone III) with smaller amounts of iron. In this study, the impact of pH on the Fe(III)-reducing activities in both of these sediment zones was investigated, and molecular analyses that elucidated the sediment microbial diversity were performed. Fe(II) was formed in zone I and III sediment microcosms at rates that were approximately 710 and 895 nmol cm−3 day−1, respectively. A shift to pH 5.3 conditions increased Fe(II) formation in zone I by a factor of 2. A shift to pH 3 conditions inhibited Fe(II) formation in zone III. Clone libraries revealed that the majority of the clones from both zones (approximately 44%) belonged to the Acidobacteria phylum. Since moderately acidophilic Acidobacteria species have the ability to oxidize Fe(II) and since Acidobacterium capsulatum reduced Fe oxides at pHs ranging from 2 to 5, this group appeared to be involved in the cycling of iron. PCR products specific for species related to Acidiphilium revealed that there were higher numbers of phylotypes related to cultured Acidiphilium or Acidisphaera species in zone III than in zone I. From the PCR products obtained for bioleaching-associated bacteria, only one phylotype with a level of similarity to Acidithiobacillus ferrooxidans of 99% was obtained. Using primer sets specific for Geobacteraceae, PCR products were obtained in higher DNA dilutions from zone III than from zone I. Phylogenetic analysis of clone libraries obtained from Fe(III)-reducing enrichment cultures grown at pH 5.5 revealed that the majority of clones were closely related to members of the Betaproteobacteria, primarily species of Thiomonas. Our results demonstrated that the upper acidic sediment was inhabited by acidophiles or moderate acidophiles which can also reduce Fe(III) under slightly acidic conditions. The majority of Fe(III) reducers inhabiting the slightly acidic sediment had only minor capacities to be active under acidic conditions.  相似文献   

11.
(Z)-4',5'-Didehydro-5'-deoxy-5'-fluoroadenosine (I), 5'-deoxy-5'-difluoroadenosine (II), and 4',5'-didehydro-5'-deoxy-5'-fluoroarabinosyl-adenosine (III) are inhibitors of rat liver S-adenosyl-L-homocysteine hydrolase. Compounds I and II are time-dependent and irreversible inhibitors of the enzyme. Both I and II are oxidized by E.NAD to produce E.NADH, and fluoride anion is formed in the inactivation reaction (0.7 to 1.0 mole fluoride/mole of enzyme subunit, and 1.7 moles fluoride/mole of enzyme subunit from I and II, respectively). The enzyme is stoichiometrically labeled with [8-3H]-I, but the label is lost upon denaturation of the protein either with or without treatment of the labeled complex with sodium borohydride. The compound III, the arabino derivative of I, is a competitive inhibitor of the enzyme. The mechanism of the inhibition of S-adenosyl-L-homocysteine hydrolase by these inhibitors is discussed.  相似文献   

12.
The rates of reduction of horseradish peroxidase compound II by p-methoxyphenol (4-hydroxyanisole) have been studied from pH 6.0 to 10.5. The kinetics are influenced by an acid group of pKa 8.7 on compound II. The acidic form of compound II is reactive; the basic form is not. Only the electrically neutral, unionized form of p-methoxyphenol is reactive. Fifteen different phenols were reacted with compound II at either pH 7.6 or pH 7.0 (three of them at both pH's). Rate constants varied from zero for p-nitrophenol to 3.2 X 10(7) M-1 for p-aminophenol. The reactive m- and p-substituted phenols yield a rho value of -4.6 +/- 0.5 when plotted according to the Hammett relation. This compares to the rho value of -6.9 obtained for horseradish peroxidase compound I reactions with phenols (1976, D. Job and H. B. Dunford, Eur. J. Biochem. 66, 607). The difference in sensitivity of compounds I and II to electron donating substituents on the phenols can be explained in terms of the relative simplicity of the reactions. Electron donation occurs to the electron-deficient porphyrin pi-cation radical of compound I accompanied by single proton addition to the protein. For compound II the electron is fed to the ferryl group at the center of the porphyrin in a reaction accompanied by two proton additions to the ferryl oxygen atom, one from the protein and the other from the substrate or solvent. This is followed by loss of water from the inner coordination sphere of the ferric ion. The relative reactivities of three o-substituted phenols can be explained in terms of steric hindrance which is minimal for a single o-substituent.  相似文献   

13.
An inhibitor of malted barley (Hordeum vulgare cv Conquest) α-amylase II was purified 125-fold from a crude extract of barley kernels by (NH4)2SO4 fractionation, ion exchange chromatography on DEAE-Sephacel, and gel filtration on Bio-Gel P 60. The inhibitor was a protein with an approximate molecular weight of 20,000 daltons and an isoelectric point of 7.3. The protein was homogeneous, as assessed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Amino acid analysis indicated the presence of about 9 half-cystine residues per mole. The neutral isoelectric point of the inhibitor suggested that some of the apparently acidic residues (glutamic and aspartic) existed in the amide form. The first twenty N-terminal amino acids were sequenced. Some homology appeared to exist between the α-amylase II inhibitor and trypsin inhibitor from barley. Complex formation between α-amylase II and the inhibitor was detected by the appearance of a new molecular weight species after gel filtration on Bio-Gel P 100. Enzyme and inhibitor had to be preincubated for 5 min, prior to assaying for enzyme activity before maximum inhibition was attained. Inhibition increased at higher pH values. At pH 5.5, an approximately 1100 molar excess of inhibitor over α-amylase II produced 40% inhibition, whereas, at pH 8.0, a 1:1 molar ratio of inhibitor to enzyme produced the same degree of inhibition.  相似文献   

14.
DPNH peroxidase is a flavin adenine dinucleotide-containing flavoprotein. Anaerobic titration of enzyme with dithionite has shown that the active site of the enzyme contains 2 mol of flavin and in addition 1 mol of a non-flavin electron acceptor that is tentatively identified as a disulfide group. Thus complete reduction of the enzyme requires 3 mol of dithionite per mole of active site. The first mole of dithionite reduces the non-flavin acceptor; complex formation between the reduced acceptor and one of the bound flavin molecules causes the formation of a long wavelength absorption band between 500 and 670 nm. The second mole of dithionite reduces the flavin that interacts with the reduced non-flavin group, and the long wavelength band disappears. The third mole of dithionite reduces the second mole of flavin. All groups are reoxidized in the presence of air. DPNH reacts with only two of the enzyme-bound electron acceptors. The first mole of DPNH reduces the non-flavin group to form an intermediate (I) that is almost identical with that formed by dithionite. The second mole of DPNH complexes with the second flavin of Intermediate I to form Intermediate II. This reaction causes a further absorbance increase in the long wavelength region; the tail of the absorption band now extends to 960 nm. The titration data (potassium phosphate, 0.05 M, pH 7.0) can be fitted with dissociation constants of 1 times 10-7 M for the formation of I, and 3 times 10-6 M for the conversion of I to II. In air, species II is oxidized to I; I is stable in air, but is oxidized stoichiometrically to oxidized enzyme by H2O2. Present evidence suggests that bound DPN-plus is responsible for the air stability of species I. Intermediate I, but not oxidized enzyme, reacts slowly with phenylmercuric acetate. This reaction causes loss of the air-stable intermediate and parallel loss in enzyme activity. The inactive enzyme cannot be reduced by DPNH to Species I; DPNH can, however, still react with the second flavin to form the autoxidizable complex. With other methods of enzyme inactivation there is also a direct correlation between residual enzyme activity and the ability of enzyme to form the air-stable intermediate. It is concluded that the air-stable intermediate is an important catalytic species.  相似文献   

15.
5'-Deiodination of thyroxine (yielding 3,3',5-tri-iodothyronine; reaction I) and of 3,3',5'-tri-iodothyronine (yielding 3,3'-di-iodothyronine; reaction II) and 5-deiodination of thyroxine (yielding 3,3',5'-tri-iodothyronine; reaction III) and of 3,3',5-tri-iodothyronine (yielding 3,3'-di-iodothyronine; reaction IV) as catalysed by rat liver microsomal fraction were studied at pH 6.5, 7.2 and 8.0 It was found that: (1) the Km of reaction I was relatively independent of pH (approx. 3 microM), whereas V was highest at pH 6.5 (63 pmol of 3,3',5-tri-iodothyronine/min per mg of protein); (2) the Km of reaction II was lowest at pH 6.5 (0.035 microM), but V was highest at pH 8.0 (829 pmol of 3,3'-di-iodothyronine/min per mg of protein); (3) thyroxine inhibited reaction II competitively; Ki values were identical at pH 6.5 and 8.0 (1 microM); (4) for both reactions III and IV Km was lowest and V was highest at pH 8.0. The results are compatible with the view that reactions I and II are mediated by a single enzyme (iodothyronine 5'-deiodinase) and that reactions III and IV are catalysed by a second enzyme (iodothyronine 5-deiodinase).  相似文献   

16.
Light-scattering and related studies on protein of Dahlmense strain of tobacco mosaic virus (DTMV) show that its polymerization characteristics are considerably different from those of TMV protein. At pH 6.0 in phosphate buffer (I = 0.1), the extent of polymerization of DTMV protein is greater than that of TMV protein, they are nearly the same at pH 6.25, and that of DTMV protein is less than that of TMV protein at pH 6.5. At pH 7.0 and 7.5, DTMV protein polymerizes more readily than TMV protein. Similar studies in phosphate buffer (I = 0.05) show that the extent of polymerization for DTMV protein is less than that of TMV protein at pH 6.0 and almost negligible at pH 6.25. Acid-base titration studies show that, upon temperature-mediated polymerization, about 2 H+ ions are bound per monomer of DTMV protein at pH 6.O.Electron microscope studies show that DTMV protein exists at room temperature as double discs and polymerized rods in phosphate buffer at pH 7.5, I = 0.1; at pH values below 6.5, DTMV protein is entirely in the form of polymerized rods. Velocity sedimentation studies of DTMV protein at room temperature are in agreement with these findings. At low temperatures, except at pH 7.5, most of the material sedimented with an s value of around 25 S. Thus, at low temperatures, except at pH 7.5, DTMV protein in solution is in the form of particles the size of double discs with an M?r of 596,000 g/mole or even larger. Therefore, temperature-mediated polymerization of DTMV protein at pH values below 6.5 in phosphate buffer (I = 0.1) and below 6.25 in phosphate buffer (I = 0.05) involves particles at least as large as double discs as the starting material.  相似文献   

17.
Enzymes I and II, which have a high soymilk-clotting activity, produced from K-295G-7 were purified by chromatographies on Sephadex G-100, CM-cellulose, hydroxylapatite, and 2nd Sephadex G-100.

The two purified enzymes were found to be homogeneous by polyacrylamide gel elec-trophoresis (PAGE) at pH 4.3. The molecular weights of enzymes I and II were 28,000 and 29,500 by SDS-PAGE, and their isoelectric points were 9.22 and 9.45, respectively. Enzymes I and II coagulated soymilk optimally at 65°C and were stable up to 45°C. Both enzymes were most active at pH 5.8, for soymilk coagulation between pH 5.8 to 6.7, and were stable with about 50 ~ 100% of the original activity from pH 5 to 10.

Each of the purified enzymes was a serine protease with an optimum pH of 9.0 for soy protein isolate (SPI) and casein digestions, because these enzymes were inhibited completely by diisopropylfluoro-phosphate (DFP).

The soymilk-clotting activity to proteolytic activity ratio of the enzyme II was 3 times higher than that of enzyme I. Enzymes I and II were more sensitive to the calcium ion concentration in soymilk than bromelain is.  相似文献   

18.
B. Hock 《Planta》1973,112(2):137-148
Summary Five MDH-isoenzymes (I–V) from cotyledons of dark-grown water-melon seedlings older than 2 days can be identified by disc-electrophoresis. By isolating and fractionating cell organelles (10000 g fraction) by density gradient centrifugation (Fig. 1) the following compartmentation of the MDH-isoenzymes can be shown: the mitochondria contain isoenzyme III and the glyoxysomes preponderantly (if not exclusive) isoenzyme V (Fig. 2), whereas the isoenzymes I, II, and IV belong to the cytosol.The 5 MDH-isoenzymes differ in several properties, e.g. differential precipitation by ammonium sulfate (Fig. 4) or isoelectric point (Fig. 5). The glyoxysomal MDH is a relatively basic protein (isoelectric point at pH 8.7), whereas the isoelectric points of the other isoenzymes lie between pH 6.4 (IV) and pH 4.7 (I).  相似文献   

19.
Disulfide bonds in soybean trypsin inhibitor (Kunitz) were simultaneously reduced and alkylated using tri-n-butylphosphine and 2-vinylquinoline at pH 7.6 in 0.11 M Tris-4.4 M urea, 41% ethanol. The resulting S-β-2-quinolylethylated protein (2-QE-STI) has a new absorption peak at 315–318 nm. Its quinoline fluorescence can be excited above 310 nm independently of intrinsic protein fluorescence. Free 2-quinolylethylcysteine (2-QEC) shows unexpectedly weak fluorescence. Quinoline absorption in 2-QEC and 2-QE-STI changes with pH. The apparentpK values determined spectrophotometrically are near 5 for 2-QEC and 3 for 2-QE-STI. Fluorescence decreased with increasing pH and in the presence of chloride ions. Both structural and charge effects thus appear to influence the absorption and fluorescence of the quinoline group. Corrected fluorescence emission (excited at 316 nm) of neutral 2-QE-STI diluted in 0.1 N H2SO4 was directly proportional to concentration in the range 0.4–8 μm 2-QEC. The 2-QEC content of the protein derivative determined by UV absorption at pH 1.5 was in agreement with the expected value of four residues per mole. Fluorescence measurements ofS-2-quinolylethylated proteins may be especially useful as a sensitive, specific assay for cyst(e)ine residues.  相似文献   

20.
P.M. Dey 《Phytochemistry》1981,20(7):1493-1496
The major sugars of fresh seeds of Castanea sativa were shown to be raffinose, stachyose and sucrose. Drying seeds at 25° for 14 weeks increased the ratio raffinose: stachyose from 1.1 to 3.5, reduced sucrose content by ca 50 % and decreased total extractable α-galactosidase. The enzyme activity was resolved into two peaks, a high MW form I (apparent MW215 000) and a low MW form II (apparent MW 53 000). The latter form was predominant in the extract of fresh seeds whereas the former was the main form in the 14-week dried seeds. An increase in the amount of enzyme I was also observed when a buffered extract (pH 5.5) of fresh seeds was stored at 4°. Enzymes I and II had pH optima of 4.5 and 6, respectively. Both enzymes hydrolysed p-nitrophenyl α-d-galactoside at a much greater rate than the natural substrates raffinose, stachyose, locust bean gum and carob gum. However, enzyme I showed preference for stachyose as compared to raffinose; the opposite order was observed for enzyme II.  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