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.
Fractionation of pectic polysaccharides from the juice of ripening `Bartlett' pears (Pyrus communis) gave two general types of polyuronides. The major type was a homogalacturonan (HGA) whose molecular weight decreased upon ripening. The other type comprised heteropolymers composed of various amounts of arabinose, rhamnose, and galactose. Treatment of the major arabinose-containing heteropolymeric fraction of high molecular weight (400,000) with a pear exo-polygalacturonase to degrade contaminating HGA gave a polyuronide which was inert to tomato endopolygalacturonase. Glycosyl-linkage analysis of this arabinosyl-polyuronide gave results expected from a rhamnogalacturonan I-like polysaccharide with large, highly branched araban side chains (RG-I). A linkage between HGA and RG-I was not found. RG-I, in ripening pears, appeared to be degraded with the initial loss of much of its arabinose.  相似文献   

2.
Pectins from sugar beet, lime and apple were degraded by a rhamnogalacturonan hydrolase associated or not with pectin methylesterases and side chain degrading enzymes (galactanase and arabinanase). The composition of the enzymatic mixture was optimised by following the reaction by viscosimetric means. The reaction products were fractionated by ion exchange chromatography. Treatment with all the enzymes released four fractions: (1). 227-247 mg/g of initial pectins and corresponded to neutral sugars from the side chains; (2,3). represented together 184-220 mg/g of pectins and corresponded to rhamnogalacturonan; (4). 533-588 mg/g of pectins and corresponded to homogalacturonan. Lime pectins have the shortest rhamnogalacturonan regions. The molar masses of homogalacturonans were in the range of 16000-43400 g/mol according to the origin of pectins, corresponding to degrees of polymerisation of 85-250. The mode of action of the enzymes used is also discussed.  相似文献   

3.
The cell-wall polysaccharides of Arabidopsis thaliana leaves have been isolated, purified, and characterized. The primary cell walls of all higher plants that have been studied contain cellulose, the three pectic polysaccharides homogalacturonan, rhamnogalacturonan I and rhamnogalacturonan II, the two hemicelluloses xyloglucan and glucuronoarabinoxylan, and structural glycoproteins. The cell walls of Arabidopsis leaves contain each of these components and no others that we could detect, and these cell walls are remarkable in that they are particularly rich in phosphate buffer-soluble polysaccharides (34% of the wall). The pectic polysaccharides of the purified cell walls consist of rhamnogalacturonan I (11%), rhamnogalacturonon II (8%), and homogalacturonan (23%). Xyloglucan (XG) accounts for 20% of the wall, and the oligosaccharide fragments generated from XG by endoglucanase consist of the typical subunits of other higher plant XGs. Glucuronoarabinoxylan (4%), cellulose (14%) and protein (14%) account for the remainder of the wall. Except for the phosphate buffer-soluble pectic polysaccharides, the polysaccharides of Arabidopsis leaf cell walls occur in proportions similar to those of other plants. The structure of the Arabidopsis cell-wall polysaccharides are typical of those of many other plants.  相似文献   

4.
Considerable information has been obtained about the primary structures of suspension-cultured sycamore (Acer pseudoplatanus) cell-wall pectic polysaccharides, i.e. rhamnogalacturonan I, rhamnogalacturonan II, and homogalacturonan. However, these polysaccharides, which are solubilized from the walls by endo-α-1,4-polygalacturonase, account for only about half of the pectic polysaccharides known to be present in sycamore cell walls. We now report that, after exhaustive treatment with endo-α-1,4-polygalacturonase, additional pectic polysaccharides were extracted from sycamore cell walls by treatment with Na2CO3 at 1 and 22°C. These previously uncharacterized polysaccharides accounted for ~4% of the cell wall. Based on the glycosyl and glycosyl-linkage compositions and the nature of the products obtained by treating the quantitatively predominant NaCO3-extracted polysaccharides with lithium metal dissolved in ethylenediamine, the polysaccharides were found to strongly resemble rhamnogalacturonan I. However, unlike rhamnogalacturonan I that characteristically had equal amounts of 2- and 2,4-linked rhamnosyl residues in its backbone, the polysaccharides extracted in Na2CO3 at 1°C had markedly disparate ratios of 2- to 2,4-linked rhamnosyl residues. We concluded that polysaccharides similar to rhamnogalacturonan I but with different degrees of branching are present in the walls of suspension-cultured sycamore cells.  相似文献   

5.
《Carbohydrate polymers》1987,7(2):143-158
A rhamnogalacturonan, extracted with hot water from the aqueous ethanol insoluble residue of flue-cured bright tobacco lamina, was purified by tangential flow ultrafiltration, ion chromatography and gel filtration. It was characterized by chemical and spectroscopic methods. Fractionation revealed that the rhamnogalacturonan consisted of a series of polysaccharides with different amounts of methyl-esterified galactopyranosyluronic acid residues in the backbone and different amounts of neutral sugar residues.The main pectic polysaccharide fraction has a backbone consisting of 4-linked α-d-galactopyranosyluronic acid residues interspersed with 2-linked l-rhamnopyranosyl residues. Approximately 22% of the galactopyranosyluronic acid residues are methylated. The main chain is branched at C-4 of rhamnose with neutral sugar side chains containing terminal and 4-linked β-d-galactopyranosyl and terminal and 5-linked α-l-arabinofuranosyl residues. The average degree of polymerization of this tobacco rahamnogalacturonan was estimated to be 400.  相似文献   

6.
Okra pods are commonly used in Asia as a vegetable, food ingredient, as well as a traditional medicine for many different purposes; for example, as diuretic agent, for treatment of dental diseases and to reduce/prevent gastric irritations. The healthy properties are suggested to originate from the high polysaccharide content of okra pods, resulting in a highly viscous solution with a slimy appearance when okra is extracted with water. In this study, we present a structural characterisation of all major cell wall polysaccharides originating from okra pods. The sequential extraction of okra cell wall material yielded fractions of soluble solids extractable using hot buffer (HBSS), chelating agent (CHSS), dilute alkaline (DASS) and concentrated alkaline (CASS). The HBSS fraction was shown to be rich in galactose, rhamnose and galacturonic acid in the ratio 1.3:1:1.3. The degree of acetylation is relatively high (DA = 58) while the degree of methyl esterification is relatively low (DM = 24). The CHSS fraction contained much higher levels of methyl esterified galacturonic acid residues (63% galacturonic acid; DM = 48) in addition to minor amounts of rhamnose and galactose. The ratio of galactose to rhamnose to galacturonic acid was 1.3:1.0:1.3 and 4.5:1.0:1.2 for HBSS and CHSS, respectively. These results indicated that the HBSS and CHSS fractions contain rhamnogalacturonan type I next to homogalacturonan, while the latter is more prevailing in CHSS. Also the DASS fraction is characterised by high amounts of rhamnose, galactose, galacturonic acid and some arabinose, indicating that rhamnogalacturonan I elements with longer arabinose- and galactose-rich side chains were part of this fraction. Partial digestion of HBSS and CHSS by pectin methyl esterase and polygalacturonase resulted in a fraction with a lower Mw and lower viscosity in solution. These samples were subjected to NMR analysis, which indicated that, in contrast to known RG I structure, the acetyl groups in HBSS are not located on the galacturonic acid residues, while for CHSS only part of the acetyl groups are located on the RG I galacturonic acid residues. The CASS fraction consisted of XXXG-type xyloglucan and 4-methylglucuronoxylan as shown by their sugar (linkage) composition and enzymatic digestion.  相似文献   

7.
The accuracy and reliability of the recently proposed scaling-relaxation method for loop closure were examined by using extensive conformational sampling. For each of the eight heptapeptides chosen to represent a variety of protein conformations, 1,000–2,000 conformations were sampled. Each segment contained 14 rotatable backbone dihedral angles. The average root mean square deviations (RMSDs) between the predicted and the native conformations were 0.7 Å for the backbone and 1.2 Å for the side chain atoms. These predictions were substantially more accurate than the previous predictions (1.1 Å for the backbone and 2.2 Å for the side chain atoms) of the same eight protein segments based on limited conformational sampling (100 conformations for each segment). Large prediction errors mostly occurred at polar and surface side chains that are unlikely to have any meaningful conformation. Moreover, the reliability of seven of the eight predictions was demonstrated with their energy-RMSD and stability-RMSD correlations of the low-energy conformations, where the conformational stability was estimated by using the multiple copy simultaneous sampling method.  相似文献   

8.
Realistic polymer chain models are developed for several polysaccharides to illustrate, using perspective drawings of representative chain conformations, the wide range of configuration, extension, and flexibility found in chains of the polysaccharide class. A method for incorporating the gauche or exo-anomeric effect into polysaccharide conformational energy functions is described, and a novel measure of directional correlation and pseudohelical persistence is utilized to help distinguish the differences in chain configuration observed among the polysaccharides compared.  相似文献   

9.
The mutant strain M5N1 CS of Rhizobium meliloti produces, in a Rhizobium complete medium supplemented with fructose and sucrose, a partially acetylated homopolymer of D -glucuronic acid residues linked β-(1 → 4). This polysaccharide forms thermoreversible gels with monovalent salts and thermally stable gels with divalent salts. In order to define the different levels of structural characterization, modeling simulations were performed for both the regular (1 → 4)-β-D -glucuronan and the acetylated derivatives. This required the evaluation of the accessible conformational space for the 16 disaccharides. Detailed conformational analysis was accomplished using the flexible residue of the MM3 molecular mechanics procedure and the results were used to access the configurational statistics of representative polysaccharide chains. Within the potential energy surfaces calculated for each disaccharide, several low energy conformers can be identified. When these conformations are extrapolated to regular polysaccharide structures, they generate polymers with right- and left-handed chirality along with a 2-fold axis. This later arrangement (n = 2, h = 5.16 Å) closely corresponds to that derived from a fiber x-ray diffraction investigation. The insertion of acetyl groups induces changes in the helical features of the polymer. As for the simulation of the configurational properties of (1 → 4)-β-D -glucuronan, an extended disordered chain having a persistence length of 105 Å (corresponding to 22 monomers) is predicted. This agrees with previous conclusions derived from solution study. The inclusion of varying amounts of acetyl groups only slightly perturbs the calculated persistence length. © 1998 John Wiley & Sons, Inc. Biopoly 45: 165–175, 1998  相似文献   

10.
Soybean soluble polysaccharides (SSPS) extracted from soybean cotyledons are acidic polysaccharides and have a pectin-like structure. The results of a structural analysis of SSPS by using polygalacturonase (PGase) and rhamnogalacturonase (RGase) clarified that the main backbone consisted of galacturonan (GN) and rhamnogalacturonan (RG), which were composed of the diglycosyl repeating unit, -4)-alpha-D-GalpA-(1-->2)-alpha-L-Rhap-(1-. The side chains of beta-1,4-galactans, branched with fucose and arabinose residues, were linked to the C-4 side of rhamnose residues in the RG regions. The degree of polymerization (dps) of GN, which linked the RG regions together, was estimated to be about 4-10 residues, and some were modified with xylose residues on the C-3 side of the galacturonates. The dps of GN at the reducing end of SSPS was estimated to be about 7-9 residues. Moreover, the fragment of the basic structure of the RG region, -[4)-alpha-D-GalpA-(1-->2)-alpha-L-Rhap-(1-]2-, some of which had long-chain beta-1,4-galactans branched on the C-4 side of rhamnose residues, were liberated from SSPS by the RGase treatment. The dps of the galactan side chain was estimated to be about 43-47 residues by an analysis of the digestion products from the beta-galactosidase treatment.  相似文献   

11.
R Shogren  T A Gerken  N Jentoft 《Biochemistry》1989,28(13):5525-5536
The effect of carbohydrate on the conformation and chain dimensions of mucous glycoproteins was investigated by using light-scattering and circular dichroism studies of native, asialo, and deglycosylated (apo) ovine submaxillary gland mucin (OSM). OSM is a large glycoprotein that is extensively O-glycosylated by the disaccharide alpha-NeuNAc(2-6)alpha-GalNAc-O-Ser/Thr. Measurements of root mean square radius of gyration, (Rg2)1/2, and hydrodynamic radius, Rh, for OSM and its derivatives were carried out as a function of molecular weight by using static and dynamic light-scattering techniques. The results were fit to the wormlike chain model for describing the dimensions of extended polymer chains. By use of this model, values of h, the length per amino acid residue, and q, the persistence length, which is a measure of chain stiffness, were obtained. These values were then used to assess the conformation and degree of chain extension of intact OSM and its partially and totally deglycosylated derivatives. Native and asialo mucin are found to be highly extended random coils, with asialo mucin having a somewhat less extended structure than intact mucin. Upon the complete removal of the carbohydrate side chains, the extended structure characteristic of intact and asialo mucin collapses to chain dimensions typical of denatured globular proteins. Conformational analyses based on the rotational isomeric state model were also performed by using the probability maps of N-acetyl-O-(GalNAc)-Thr-N-methylamide as starting conformations for native and asialo mucin. The results suggest that both the glycosylated and nonglycosylated residues in native mucin may occupy a small region of conformational space having -90 degrees less than phi less than -60 degrees and 60 degrees less than psi less than 180 degrees, while a slightly broader range is found to fit asialo mucin. The proposed conformations obtained for these mucins are consistent with their circular dichroism spectra. Significantly larger ranges of phi and psi values were obtained for apo mucin, as would be expected from its circular dichroism spectra and increased flexibility. These results indicate the expanded mucin structure is the direct result of peptide core glycosylation. These observations together with the results of earlier studies indicate that steric interactions of the O-linked GalNAc residue with the peptide core are primarily responsible for the expanded mucin structure and that these perturbations extend to the nonglycosylated amino acid residues. This expanded mucin conformation must be a significant determinant of the viscoelastic properties of these molecules in solution.  相似文献   

12.
Seven differently linked glycosyl residues have been found to be glycosidically linked to O-4 of the branched 2,4-linked l-rhamnosyl residues contained in the rhamnosyl and galacturonosyl backbone of the cell wall pectic polysaccharide rhamnogalacturonan I. These seven glycosyl residues are, therefore, the first residues of at least seven different side chains attached to the rhamnogalacturonan backbone. These first side chain glycosyl residues are 5-linked l-arabinofuranosyl and terminal 3-, 4-, 6-, 2,6-, and 3,6-linked d-galactopyranosyl residues. The existence of at least seven different side chains in rhamnogalacturonan I indicates that rhamnogalacturonan I is either an exceedingly complex polysaccharide or that rhamnogalacturonan I is a family of polysaccharides with similar or identical rhamnogalacturonan backbones substituted with different side chains.  相似文献   

13.
14.
Silenan SV, a pectic polysaccharide, was isolated from the aerial part of Silene vulgaris (Moench) Garke (Oberna behen (L.) Ikonn.), widespread through the European North of Russia. The polysaccharide was found to contain residues of galacturonic acid (63%), arabinose, galactose, and rhamnose as the main constituents. The results of a partial acidic hydrolysis, pectinase digestion, and NMR studies of silenan SV indicated that its molecule contains a linear alpha-1,4-D-galacturonan backbone and ramified regions. The core of the ramified regions is composed of residues of alpha-1,4-D-galacturonic acid along with 2-substituted alpha-rhamnopyranose residues. The NMR data showed that the silenan SV side chains are composed of the blocks built from the terminal alpha-1,5-linked arabinofuranose and beta-1,4-linked galactopyranose residues; these most likely are the side chains of rhamnogalacturonan, characteristic of other pectic polysaccharides. The nonreducing ends of these side chains contain alpha-arabinofuranose residues.  相似文献   

15.
The okra plant, Abelmoschus esculentus (L.) Moench, a native plant from Africa, is now cultivated in many other areas such as Asia, Africa, Middle East, and the southern states of the USA. Okra pods are used as vegetables and as traditional medicines. Sequential extraction showed that the Hot Buffer Soluble Solids (HBSS) extract of okra consists of highly branched rhamnogalacturonan (RG) I containing high levels of acetyl groups and short galactose side chains. In contrast, the CHelating agent Soluble Solids (CHSS) extract contained pectin with less RG I regions and slightly longer galactose side chains. Both pectic populations were incubated with homogeneous and well characterized rhamnogalacturonan hydrolase (RGH), endo-polygalacturonase (PG), and endo-galactanase (endo-Gal), monitoring both high and low molecular weight fragments. RGH is able to degrade saponified HBSS and, to some extent, also non-saponified HBSS, while PG and endo-Gal are hardly able to degrade either HBSS or saponified HBSS. In contrast, PG is successful in degrading CHSS, while RGH and endo-Gal are hardly able to degrade the CHSS structure. These results point to a much higher homogalacturonan (HG) ratio for CHSS when compared to HBSS. In addition, the CHSS contained slightly longer galactan side chains within its RG I region than HBSS. Matrix-assisted laser desorption ionization-time of flight mass spectrometry indicated the presence of acetylated RG oligomers in the HBSS and CHSS enzyme digests and electron spray ionization-ion trap-mass spectrum showed that not only galacturonosyl residues but also rhamnosyl residues in RG I oligomers were O-acetylated. NMR spectroscopy showed that all rhamnose residues in a 20 kDa HBSS population were O-acetylated at position O-3. Surprisingly, the NMR data also showed that terminal α-linked galactosyl groups were present as neutral side chain substituents. Taken together, these results demonstrate that okra contained RG I structures which have not been reported before for pectic RG I.  相似文献   

16.
K. H. Gardner  J. Blackwell 《Biopolymers》1974,13(10):1975-2001
Native cellulose has been shown to consist of a crystalline array of parallel chains, based on the X-ray diffraction data for specimens from the sea alga Valonia ventricosa. The unit cell is monoclinic with dimensions a = 16.34 Å, b = 15.72 Å, c = 10.38 Å (fiber axis), and β = 97.0°. The space group is P21 and the cell contains disaccharide segments of eight chains. Models containing chains with the same sense (parallel) or alternating sense (antiparallel) were refined against the intensity data using rigidbody least squares procedures. The results show a preference for a parallel chain structure with specific chain polarity with respect to the c axis. The refinement places the ? CH2OH side chains approximately 20′ from the so-called tg conformation, with a result that an 02′? H…06 intramolecular bond is formed. The structure also contains an 03? H…05′ intramolecular bond and an 06? H…03 intermolecular bond along the a axis. All these bonds lie in the 020 planes, and the structure is an array of hydrogen-bonded sheets. A major consequence of this work is that regular chain folding can be ruled out and cellulose is seen as extended chain polymer single crystals.  相似文献   

17.
The covalent attachment of different types of poly‐ubiquitin chains signal different outcomes for the proteins so targeted. For example, a protein modified with Lys‐48‐linked poly‐ubiquitin chains is targeted for proteasomal degradation, whereas Lys‐63‐linked chains encode nondegradative signals. The structural features that enable these different types of chains to encode different signals have not yet been fully elucidated. We report here the X‐ray crystal structures of Lys‐63‐linked tri‐ and di‐ubiquitin at resolutions of 2.3 and 1.9 Å, respectively. The tri‐ and di‐ubiquitin species adopt essentially identical structures. In both instances, the ubiquitin chain assumes a highly extended conformation with a left‐handed helical twist; the helical chain contains four ubiquitin monomers per turn and has a repeat length of ~110 Å. Interestingly, Lys‐48 ubiquitin chains also adopt a left‐handed helical structure with a similar repeat length. However, the Lys‐63 architecture is much more open than that of Lys‐48 chains and exposes much more of the ubiquitin surface for potential recognition events. These new crystal structures are consistent with the results of solution studies of Lys‐63 chain conformation, and reveal the structural basis for differential recognition of Lys‐63 versus Lys‐48 chains. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

18.
His I  Driouich A  Nicol F  Jauneau A  Höfte H 《Planta》2001,212(3):348-358
Korrigan (kor) is a dwarf mutant of Arabidopsis thaliana (L.) Heynh. that is deficient in a membrane-bound endo-1,4-beta-glucanase. The effect of the mutation on the pectin network has been studied in kor by microscopical techniques associated with various probes specific for different classes of pectic polysaccharides. The localisation of native crystalline cellulose was also examined using the cellobiohydrolase I-gold probe. The investigations were focused on the external cell walls of the epidermis, a cell layer that, in a number of plant species, has been shown to be growth limiting. Anionic sites associated with pectic polymers were quantified using the cationic gold probe. Homogalacturonans were quantified using polyclonal anti-polygalacturonic acid/rhamnogalacturonan I antibodies recognising polygalacturonic acid, and monoclonal JIM7 and JIM5 antibodies recognising homogalacturonans with a high or low degree of methyl-esterification, respectively. Rhamnogalacturonans were quantified with two monoclonal antibodies, LM5, recognising beta-1,4 galactan side chains of rhamnogalacturonan I, and CCRCM2. Our results show a marked increase in homogalacturonan epitopes and a decrease in rhamnogalacturonan epitopes in kor compared to the wild type. A substantial decrease in cellobiohydrolase I-gold labelling was also observed in the mutant cell walls. These findings demonstrate that a deficiency in an endo-1,4-beta-glucanase, which is in principle not directly implicated in pectin metabolism, can induce important changes in pectin composition in the primary cell wall. The changes indicate the existence of feedback mechanisms controlling the synthesis and/or deposition of pectic polysaccharides in primary cell walls.  相似文献   

19.
Individual pectin polymers and complexes, isolated from the pericarp of unripe tomato (Lycopersicon esculentum var. Rutgers), were subjected to a mild acid hydrolysis and visualised and characterised by atomic force microscopy (AFM). The AFM images confirm earlier studies showing that individual pectic polysaccharides often possess long branches. The AFM data have been used to construct size and molecular weight distributions for the single molecules and complexes, from which the calculated number-average and weight-average molecular weights can then be compared directly with the published literature data on the rheology of bulk samples. Loss of the neutral sugars arabinose, galactose and rhamnose from the pectin samples does not significantly alter either the size or the branching density of the individual polymers, but is reflected in a breakdown of the complexes. Significant loss of galacturonic acid at long hydrolysis times was found to be accompanied by changes in the size and branching of the single polymers and further breakdown of the complexes. The results suggest that rhamnose, arabinose and galactose are not the major components of the individual polymers but are, instead, confined to the complexes. The polysaccharides represent a previously unrecognised branched homogalacturonan with a minimum mean size some three times larger than that previously reported. The complexes consist of homogalacturonans (HGs) held together by rhamnogalacturonan I (RG-I) regions. Comparison of the rate of depolymerisation of the homogalacturonans and complexes with the published data on changes in the intrinsic viscosity of bulk pectin samples, subjected to similar acid hydrolysis, suggests that the different rates of depolymerisation of RG-I and HG contribute separately to the observed changes in intrinsic viscosity during acid hydrolysis. Thus data obtained using a single molecule microscopy technique provides new insights into the behaviour in the bulk.  相似文献   

20.
The allosteric enzyme aspartate carbamoyltransferase of Escherichia coli consists of six regulatory chains (R) and six catalytic chains (C) in D3 symmetry. The less active T conformation, complexed to the allosteric inhibitor CTP has been refined to 2.6 A (R-factor of 0.155). We now report refinement of the more active R conformation, complexed to the bisubstrate analog N-phosphonacetyl-L-aspartate (PALA) to 2.4 A (R-factor of 0.165, root-mean-square deviations from ideal bond distances and angles of 0.013 A and 2.2 degrees, respectively). The antiparallel beta-sheet in the revised segment 8-65 of the regulatory chain of the T conformation is confirmed in the R conformation, as is also the interchange of alanine 1 with the side-chain of asparagine 2 in the catalytic chain. The crystallographic asymmetric unit containing one-third of the molecule (C2R2) includes 925 sites for water molecules, and seven side-chains in alternative conformations. The gross conformational changes of the T to R transition are confirmed, including the elongation of the molecule along its threefold axis by 12 A, the relative reorientation of the catalytic trimers C3 by 10 degrees, and the rotation of the regulatory dimers R2 about the molecular twofold axis by 15 degrees. No changes occur in secondary structure. Essentially rigid-body transformations account for the movement of the four domains of each catalytic-regulatory unit; these include the allosteric effector domain, the equatorial (aspartate) domain, and the combination of the polar (carbamyl phosphate) and zinc domain, which moves as a rigid unit. However, interfaces change, for example the interface between the zinc domain of the R chain and the equatorial domain of the C chain, is nearly absent in the T state, but becomes extensive in the R state of the enzyme; also one catalytic-regulatory interface (C1-R4) of the T state disappears in the more active R state of the enzyme. Segments 50-55, 77-86 and 231-246 of the catalytic chain and segments 51-55, 67-72 and 150-153 of the regulatory chain show conformational changes that go beyond the rigid-body movement of their corresponding domains. The localized conformational changes in the catalytic chain all derive from the interactions of the enzyme with the inhibitor PALA; these changes may be important for the catalytic mechanism. The conformation changes in segments 67-72 and 150-153 of the regulatory chain may be important for the allosteric control of substrate binding. On the basis of the conformational differences of the T and R states of the enzyme, we present a plausible scheme for catalysis that assumes the ordered binding of substrates and the ordered release o  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