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 prolamin fraction of the rye cultivar Danko was reduced with dithioerythritol and separated by reversed‐phase HPLC on C18 silica gel. Two major γ‐75k secalins, P1 and P2, were collected, purified by rechromatography, derivatized with 4‐vinylpyridine, and digested in parallel with α‐chymotrypsin, thermolysin, and trypsin. The different enzymatic hydrolyzates were preparatively separated by two‐step reversed‐phase HPLC on C18 silica gel, and the resulting peptides were characterized by sequence analysis and, in parts, by mass spectrometry. By means of overlapping peptides and by comparison with a known DNA sequence of a γ‐75k secalin (gSec2A) derived from a wheat translocation line (2RS. 2BL), 84% of the P1 sequence and 35% of the P2 sequence could be assigned. Heterogeneity at several sequence positions demonstrated that both protein preparations were not pure and contained at least two or three components. The sequence of the C‐terminal domain of P1 was almost completely determined except for one of the 148 residues which could not be identified. The partially determined sequences of P2 were highly homologous with those of P1. The results revealed a close relationship between P1, P2, and gSec2A and a high degree of homology with γ‐gliadins of wheat including eight cysteine residues in homologous positions. The partially sequenced N‐terminal domain of P1 was similar to that of gSec2A and consisted of repetitive sequences rich in glutamine, proline, and aromatic amino acids. Differences from γ‐gliadins were found in the strongly increased number of residues, in the more frequent modifications of the repetitive motifs, and in the presence of a cysteine residue at position 12. The partial amino acid sequence of the N‐terminal domain of P2 was in agreement with that of P1, besides a few exceptions in single positions and in the presence of a second cysteine residue.  相似文献   

2.
Kernels of the rye cultivars Danko and Halo were milled into white flour and compared with flour of the wheat cultivar Rektor. Flour proteins were extracted stepwise with a salt solution (albumins‐globulins), 60% ethanol (prolamins), and 50% 2‐propanol under reducing conditions (glutelins). The quantification by reversed‐phase HPLC indicated that the extractable proteins of both rye flours consisted of ≈26% albumins‐globulins, 65% prolamins, and 9% glutelins. Compared with wheat flour, rye flours comprised significantly higher proportions of nonstorage proteins (albumins‐globulins) and lower proportions of polymerized storage proteins (glutelins). SDS‐PAGE revealed that the prolamin fractions of rye contained all four storage protein types (HMW, γ‐75k, ω, and γ‐40k secalins), whereas the glutelin fractions contained only HMW and γ‐75k secalins. The quantification of secalin types by RP‐HPLC showed a close relationship between the two cultivars.The γ‐75k secalins contributed nearly half (≈46%) of the total storage proteins, followed by γ‐40k secalins (24%) and ω secalins (17%); HMW secalins (≈7%) were minor components, and 6% of eluted proteins were not identified. The amino acid composition of γ‐40k secalins corresponded to those of γ‐gliadins of wheat, whereas γ‐75k secalins were characterized by higher contents of glutamine and proline. Matrix‐assisted laser desorption/ionization and time of flight mass spectrometry (MALDI‐TOF MS) indicated molecular masses of about 52,000 (γ‐75k) and 32,000 (γ‐40k), respectively. N‐terminal amino acid sequences were homologous with those of wheat γ‐ gliadins except for position 5 (asparagine in γ‐75k and glutamine in γ‐40k secalins) and position 12 (cysteine in γ‐75k secalins). The N‐terminal amino acid sequences of HMW and ω‐secalins were homologous with those of the corresponding protein types of wheat. Gel‐permeation HPLC of prolamin fractions revealed that rye flours contained a significantly higher proportion of ethanol‐soluble oligomeric proteins than wheat flour.  相似文献   

3.
ω‐Gliadins were purified from wheat (Triticum aestivum L. ‘Butte’) flour and characterized. Although reversed‐phase HPLC (RP‐HPLC) separated the 1B‐encoded ω‐gliadins into two fractions, 1B1 and 1B2, these fractions had nearly identical amino acid compositions, three similar protein bands in SDS‐PAGE, 10 similar spots in two‐dimensional PAGE, and two similar N‐terminal amino acid sequences. The main components had a range in mass of 48,900–51,500 when estimated by mass spectrometry, significantly less than the mass estimated by SDS‐PAGE. The 1B fractions were digested with thermolysin, the peptides were separated by RP‐HPLC, the peptide mass was determined, and nine peptides from each fraction were sequenced with nearly identical results for the 1B1 and 1B2 digests. A possible consensus sequence of the 1B‐encoded ω‐gliadin internal repeat was QQQXP, where X was F, I, or L in descending order of occurrence. The 1D‐encoded ω‐gliadins were purified by RP‐HPLC as a single fraction that had one band in SDS‐PAGE, two spots in two‐dimensional PAGE, two components with mass of 41,923 and 42,770 detected by mass spectrometry, and two N‐terminal sequences. Circular dichroism (CD) spectra for the 1B and 1D ω‐gliadins were similar and were suggestive of mainly flexible random structure with a significant amount of the left‐handed polyproline II helical conformation in the 1D components.  相似文献   

4.
Semidry electroblotting is convenient and allows a rapid and efficient protein transfer from two‐dimensional polyacrylamide gel electrophoresis (2D‐PAGE) gels onto sequencer stable supports for protein microsequence analysis in a gas‐phase sequencer. Using this technique, I determined the amino acid sequences of the endosperm proteins in Japanese hexaploid commercial wheats (Triticum aestivum). Based on sequence determination of the Japanese hexaploid wheats, the endosperm protein could be easily characterized. Wheat endosperm protein, extracted in the presence of 2‐mercaptoethanol and SDS, fractionated into many protein polypeptides using 2D‐PAGE under dissociating conditions. These components were grouped into HMW glutenin subunits, α‐, β‐ or γ‐gliadins, and novel protein polypeptides by using the N‐terminal amino acid sequences. The novel endosperm protein polypeptides were detected, and two new types of N‐terminal amino acid sequences have been found for protein poly‐peptides. These polypeptides have much faster electrophoresis mobility during 2D‐PAGE and are therefore probably a much smaller size than any other peptides of endosperm protein groups found in hexaploid wheat. Ten protein polypeptides have been purified from cultivars of Japanese wheat. Some differences in the contents of amino acids for four protein polypeptide spots were apparent in Japanese wheat.  相似文献   

5.
Rye sourdough (RSD) gives rye bread mildly acidic taste and desired flavor. Flavor precursors (amino acids and small peptides) are generated in the proteolytic breakdown of rye proteins. Our aim was to study the protein degradation during RSD fermentations. Two sourdoughs were prepared of flours derived from two rye cultivars (Amilo and Akusti). RSD samples were collected during fermentations. Three protein fractions were obtained by sequential protein extraction and these were analyzed by SDS‐PAGE. Free amino nitrogen (FAN) was measured with a ninhydrin method. In addition, two rye incubations without starter microorganisms (with antibiotics) were made at pH 3.6 and 6.1, and proteinase profiles of the rye cultivars were analyzed at pH 4.3. SDS‐PAGE analysis showed that during RSD fermentations, rye proteins, especially the alcohol‐soluble secalins, were degraded. Secalins also evidently degraded during the incubation without starter microorganisms at pH 3.6. Aspartic proteinases were in the major proteinase group in both rye cultivars. This study confirms that endogenous proteinases of rye, mainly aspartic proteinases, hydrolyze rye proteins, especially secalins, during RSD fermentation. Protein degradation in rye sourdoughs may thus be enhanced by selecting rye flours with high proteolytic activity toward secalins.  相似文献   

6.
Flours from nonsprouted (ns) kernels and dried sprouted (s) kernels of transgenic rye expressing HMW glutenin subunits (HMW‐GS) 1Dy10 (L10) or 1Dx5+1Dy10 (L5+10) from wheat were compared with flours from the corresponding wildtype rye (Lwt). The crude protein content of nonsprouted flours ranged from 9.2% (Lwt) to 10.4% (L5+10) and was lowered by ≈1% due to sprouting. Flour proteins were separated into albumins/globulins, prolamins, and glutelin subunits by a modified Osborne fractionation and into SDS‐soluble and insoluble fractions. Portions of the prolamin fractions were reduced in the same manner as glutelins. The different fractions were then characterized and quantified by RP‐HPLC on C8 silica gel. The proportion of albumins/globulins did not significantly differ between transgenic lines and wildtype. The proportions of alcohol‐insoluble glutelins and SDS‐insoluble proteins drastically increased in transgenic rye due to a shift of HMW and γ‐75k secalins into the polymeric fractions. Significant differences in the proportion of highly polymeric proteins between nonsprouted and sprouted flours could not be detected. The quantitative data demonstrated that the expression of HMW‐GS led to a higher degree of polymerization of storage proteins in rye flour. The HMW‐GS combination 1Dx5+1Dy10 showed stronger effects than 1Dy10 alone. The analyzed flours contained two HMW secalins (R1, R2), whose amino acid compositions were closely related to those of 1Dy10 and 1Dx5, respectively. The amounts of R1 in Lwt flours determined by RP‐HPLC were 221 mg (ns) and 186 mg (s) per 100 g and those of R2 were 344 mg (ns) and 298 mg (s), respectively. These amounts increased to 240 mg (ns)/201 mg (s) (R1) and 479 mg (ns)/432 mg (s) (R2) in L10 flours. In L5+10 flours, the amount of R1 decreased to 150 mg (ns)/132 mg (s) while R2 increased to 432 mg (ns)/338 mg (s). The amount of HMW‐GS 1Dy10 was almost the same as that of R2 in L10 flours but was strongly increased in L5+10 flour (633 mg [ns]/538 mg [s]). HMW‐GS 1Dx5 was, by far, the major subunit in L5+10 flours (987 mg 7[ns]/896 mg [s]). The summarized amounts of all HMW subunits increased from ≈0.5 g (Lwt) to ≈1.1 g (L10) and ≈2.0 g (L5+10). Thus only L10 flours were similar to wheat flours in HMW subunit content. The baking performance of L10 flour determined by a microbaking test was improved compared with Lwt flour, whereas L5x10 flour showed very poor properties obviously due to the strongly increased proportion of highly cross‐linked glutelins. The breadmaking quality of flours from 1Dy10 seeds and wildtype seeds was reduced by the same degree when flours from sprouted seeds were analyzed.  相似文献   

7.
The aim of this study was to isolate high‐molecular‐weight (HMW) gliadins from wheat flour and to characterize the protein components that contribute to HMW gliadins. Wheat flour Akteur was extracted with a modified Osborne procedure, and the fraction soluble in 60% ethanol (total gliadins) was separated by gel‐permeation HPLC, yielding three fractions, GP1–GP3. GP1 (21.5%) consisted of oligomeric HMW gliadins, GP2 (15.2%) of ω5‐gliadins, and GP3 (63.3%) of ω1,2‐, α‐, and γ‐gliadins. Two‐dimensional SDS‐PAGE of HMW gliadins showed that interchain disulfide bonds were present in HMW gliadins. The molecular mass distribution of HMW gliadins determined by gel‐permeation HPLC was in a range from 66,000 to 680,000 with an average degree of polymerization of 13. Reduced HMW gliadins were further separated by preparative reversed‐phase HPLC into four subfractions (RP1, RP2, RP3, and RP4), which were characterized by SDS‐PAGE and semiquantitative N‐terminal sequencing. HMW gliadins of the wheat flour Akteur contained all types of gluten proteins: 48% low‐molecular‐weight glutenin subunits, 18% γ‐gliadins, 13% α‐gliadins, 9% ω1,2‐gliadins, 8% HMW glutenin subunits, and 4% ω5‐gliadins. We postulate that the existence of HMW gliadins can be explained by the presence of terminators, which interrupt the polymerization of glutenin subunits during biosynthesis and lead to polymers of limited size (oligomers) that are still soluble in aqueous ethanol.  相似文献   

8.
The enzyme transglutaminase (TG) is known to have beneficial effects on breadmaking. However, only limited information is available on the structural changes of gluten proteins caused by TG treatment. The effect of TG has, therefore, been systematically studied by means of model peptides, suspensions of wheat flours and doughs. The treatment of synthetic peptides mimicking amino acid sequences of HMW subunits of glutenin with TG results in isopeptide bonds between glutamine and lysine residues. To study the effect on gluten proteins, different amounts of TG (0 to 900 mg enzyme protein per kg) were dissolved in a buffer and added to wheat flour. The flour suspensions were incubated and centrifuged and the residues were successively extracted with water, a salt solution, 60% aqueous ethanol (gliadin fraction) and SDS solution including a reducing agent (glutenin fraction). The characterization of the fractions by amino acid analysis, SDS‐PAGE, gel permeation HPLC and reversed‐phase HPLC has indicated that the quantity of extractable gliadins decreases by increasing TG amounts. Among gliadins, the ω5‐type was affected to the greatest extent by the reduction of extractability, followed by the ω1,2‐, α‐ and γ‐types. The oligomeric portion of the gliadin fractions (HMW gliadin) was strongly reduced when flour was treated with 450 and 900 mg TG per kg of flour, respectively. In the first instance, the quantity of the glutenin fractions increased by the treatment of flour with 90 and 450 mg TG per kg of flour, and significantly decreased by the treatment of flour with 900 mg TG per kg of flour. Parallel to an increase in TG concentration, the amounts of glutenin‐bound ω‐gliadins and HMW subunits were strongly reduced, whereas the LMW subunits reached a maximal amount after treatment with 450 mg TG per kg of flour. The insoluble residue was almost free of protein when flour was treated with lower amounts of TG. Higher amounts led to a great increase of protein in the residues. The effects of TG on doughs were similar to those of flour suspensions, but less strongly pronounced probably due to the lower water content of the dough system. Sequence analysis of peptides from a thermolytic digest of the insoluble residue revealed that HMW subunits of glutenin and α‐gliadins were predominantly involved in cross‐links formed by TG treatment.  相似文献   

9.
The reactions of oxidized glutathione generated from endogenous glutathione by the addition of ascorbic acid (AA) prior to dough mixing on free thiol groups of gluten proteins have been investigated. A small amount of (35)S-labeled glutathione was added as a tracer to identify the reaction products of GSSG and free protein thiols by radioactivity measurement. First, gluten was isolated from the dough, then the gliadins were extracted, and residual glutenin was partially hydrolyzed with thermolysin. After preseparation by gel permeation chromatography, the fractions with the highest radioactivity were separated by high-performance liquid chromatography. Radioactive peptides were identified, isolated, sequenced, and assigned to amino acid sequences of gluten protein components. The isolated peptides contained exclusively the cysteine residues C(b) and C(x) of low molecular weight subunits of glutenin, which are supposed to be highly reactive in forming intermolecular disulfide bonds. From these results it can be assumed that the cysteine residues C(b) and C(x) of the low molecular weight subunits of glutenin are at least partly present in the thiol form in flour. During dough mixing they are converted to protein-protein disulfides or glutathione-protein mixed disulfides by thiol/disulfide interchange reactions. Oxidized glutathione necessary for this reaction is generated from glutathione by the action of AA. These results are in accordance with the major hypothesis about the mechanism of action of AA.  相似文献   

10.
Celiac disease (CD) is an inflammatory disorder of the upper small intestine triggered by the ingestion of wheat, rye, barley, and possibly oat products. The clinical feature of CD is characterized by a flat intestinal mucosa with the absence of normal villi, resulting in a generalized malabsorption of nutrients. The prevalence of CD among Caucasians is now thought to be in a range of 1:100–300. There is a strong genetic association with human leukocyte antigens (HLA‐)DQ2 and DQ8 and currently unknown non‐HLA genes. During the last decade, intense biochemical studies have contributed to substantial progress in understanding the general principles that determine the pathogenesis of CD. The precipitating factors of toxic cereals are the storage proteins, termed gluten in the field of CD (gliadins and glutenins of wheat, secalins of rye, and hordeins of barley). There is still disagreement about the toxicity of oat avenins. The structural features unique to all CD toxic proteins are sequence domains rich in Gln and Pro. The high Pro content renders these proteins resistant to complete proteolytic digestion by gastrointestinal enzymes. Consequently, large Pro‐ and Gln‐rich peptides are cumulated in the small intestine and reach the subepithelial lymphatic tissue. Depending on the amino acid sequences, these peptides can induce two different immune responses. The rapid innate response is characterized by the secretion of the cytokine interleukin‐15 and the massive increase of intraepithelial lymphocytes. The slower adaptive response includes the binding of gluten peptides (native or partially deamidated by tissue transglutaminase) to HLA‐DQ2 or ‐DQ8 of antigen presenting cells and the subsequent stimulation of T‐cells accompanied by the release of proinflammatory cytokines such as interferon‐γ and the activation of matrix metalloproteinases. Both immune responses result in mucosal destruction and epithelial apoptosis. Additionally, stimulated T‐cells activate B‐cells that produce serum IgA and IgG antibodies against gluten proteins (antigen) and tissue transglutaminase (autoantigen). These antibodies can be used for noninvasive screening tests to diagnose CD. The current essential therapy of CD is a strict lifelong adherence to gluten‐free diet. Dietetic gluten‐free foods produced for CD patients underlie the regulations of the Codex Alimentarius Standard for Gluten‐Free Foods. The “Draft Revised Codex Standard” edited in March 2006 proposes a maximum level of 20 mg of gluten/kg for naturally gluten‐free foods (e.g., based on rice or corn flour) and 200 mg/kg for foods rendered gluten‐free (e.g., wheat starch). Numerous analytical methods for gluten determination have been developed, mostly based on immunochemical assays, mass spectrometry, or polymerase chain reaction. So far, only two enzyme‐linked immunosorbent assays have been successfully ring‐tested and are commercially available. During the last decade, future strategies for prevention and treatment of CD have been proposed. They are based on the removal of toxic epitopes by enzymatic degradation or gene engineering and on blocking parts of the immune system. However, any alternative treatment should have a safety profile competitive with gluten‐free diet.  相似文献   

11.
Wheat presents an important genetic diversity that could be useful to look for cultivars with reduced allergencity. omega5-Gliadins have been described as major allergens for wheat allergic patients suffering from wheat-dependent exercise-induced anaphylaxis (WDEIA) and some cases of chronic urticaria (U). Our objective was to study the influence of genetic variability at the Gli-B1 locus encoding for omega5-gliadins on the reactivity of IgE antibodies from these patients. We selected cultivars expressing 13 alleles at Gli-B1 including a wheat/rye translocation and studied the reactivity to gliadins of a rabbit antiserum specific for omega5-gliadins and of IgE from 10 patients. The antiserum and IgE from nine patients with WDEIA and U strongly detected omega5-gliadins expressed by most of the Gli-B1 alleles but showed no or faint responses to the gliadins and secalins extracted from the translocated wheat. The selection of genotypes lacking the Gli-B1 locus may reduce wheat allergenicity.  相似文献   

12.
This study investigated the antifreeze activity (AF) of γ‐polyglutamic acid (γ‐PGA), freezing resistance of yeast cells and sweet dough, and the mechanism influenced by γ‐PGA. Properties studied included AF of γ‐PGA, water‐holding capacity of flour, survival ratio and oxidation resistance capability of yeast cells, ice melting enthalpy (ΔH), and fermentation and breadmaking properties of sweet dough. The AF of γ‐PGA was 8.03 g of unfrozen water/g of sample, indicating good AF. γ‐PGA was tested on yeast cells and sweet dough stored frozen for 0, 1, 2, 4, and 8 weeks at four levels (0, 0.5, 1, and 3%). Survival ratio of yeast cells with γ‐PGA was significantly higher than the corresponding control. A possible mechanism might be related to the modulation of oxidation resistance capability of yeast cells by γ‐PGA. A decrease in glutathione release from frozen yeast cells and an increase in water‐holding capacity of wheat dough were observed with the addition of γ‐PGA. In the presence of γ‐PGA, ΔH, ice melting temperature, and proofing time of frozen sweet dough decreased significantly, and fermentation parameters improved, compared with the corresponding control sample. Specific volume of bread made from frozen sweet dough with 0.5, 1, and 3% γ‐PGA increased by 6.3, 8.9, and 3.3%, respectively, after 8 weeks of frozen storage. γ‐PGA enhanced the freezing resistance of yeast cells and sweet dough effectively, and the effect on specific volume of bread was not linear, with 1% showing better results.  相似文献   

13.
A mouse lymphatic endothelial cell (SVEC4‐10) in vitro model was developed and found to be effective in the study of antioxidant activity of γ‐oryzanol in rice bran. The critical and vital parameters in developing these cell models included the emulsion preparation of hydrophobic compounds, the consistent management of cell culture, and the selection of cell viability detection methods compatible with the cell lines and the test substances. The SVEC4‐10 cell line had a fast metabolism and consequently could be used to determine antioxidant activity of a test substance in a relatively rapid manner. Results showed that γ‐oryzanol, ferulic acid, and α‐tocopherol could interact with cells such that oxidative damage induced by tert‐butyl hydroperoxide on cellular mitochondrial activity lessened, and in some situations, γ‐oryzanol was a more effective antioxidant than α‐tocopherol. The results also suggested different antioxidant mechanisms among γ‐oryzanol, ferulic acid, and α‐tocopherol. The three major components of γ‐oryzanol generally had higher antioxidant activity than γ‐oryzanol, with 24‐methylene cycloartanyl ferulate relatively more effective. Synergistic antioxidant activity among γ‐oryzanol, ferulic acid, and α‐tocopherol was also found.  相似文献   

14.
Twenty‐two lines of emmer (T. dicoccon Schrank) and 10 of spelt (T. spelta L.) were analyzed using capillary electrophoresis for their gliadins. These proteins were separated on an uncoated fused‐silica capillary (30 cm long, 22 cm to detector, 50 μm i.d.) using the isoelectric buffer 40 mM aspartic acid, 4M urea, 0.5% (w/v) HEC, and 20% (v/v) acetonitrile. Samples were run for 20 min at 22kV and 42°C. By using these conditions, gliadins were separated into 21–30 components (peaks and shoulders). The major peaks eluted between 4.5 and 8.5 min. Electrophoregrams of tested lines showed qualitative and quantitative differences, including number of peaks, presence or absence of some major peaks, and areas of peaks. Lines belonging to the same species can be discriminated mainly on the basis of β‐ and ω‐gliadin patterns. The γ‐and ω‐gliadins seem to be more useful in the differentiation of emmer from spelt. The comparison of electrophoregrams relative to hulled and unhulled species evidenced the high similarity between species with the same genome composition (durum wheat‐emmer, and common wheat‐spelt).  相似文献   

15.
Antibodies specific for wheat proteins were used to identify protein fractions modified during extrusion of Hard Red Spring wheat flour (14% protein) under four different combinations of extrusion conditions (18 and 24% feed moisture and 145 and 175°C die temperature). Antibody binding was assessed on immunoblots of proteins extracted from flour and extrudates separated by SDS‐PAGE. Antibodies to high molecular weight glutenin subunits (HMW‐GS) and to B‐group low molecular weight glutenin subunits (LMW‐GS) recognized intact subunits from both flour and extrudates. Antibodies to C‐group LMW‐GS had diminished binding to extruded proteins. Glutenin‐specific antibodies also recognized protein in the extrudates migrating as a smear at molecular weights higher than intact subunits, indicating cross‐linked proteins. Antibodies recognized albumins or globulins in flour but not in extrudates, evidence that these fractions undergo significant modification during extrusion. Acid‐PAGE and antibody reaction of gliadins extracted in 1M urea and in 70% ethanol revealed total loss of cysteine‐containing α, β, γ‐gliadins but no obvious effects on sulfur‐poor ω‐gliadins, suggesting gliadin modification involves replacing intramolecular disulfides with intermolecular disulfide cross‐links. Identifying protein fractions modified during different extrusion conditions may provide new options for tailoring extrusion to achieve specific textural characteristics.  相似文献   

16.
The complete amino acid sequence of an α-type gliadin from spelt wheat (spelta) has been deduced from the cloned DNA sequence and compared with α-type gliadin sequences from bread wheat. The comparison showed only minor differences in amino acid sequences between the α-type gliadin from bread wheat and the α-type gliadin from spelta. The two sequences had an identity of 98.5%. Larger differences can be found between different α-type gliadin amino acid sequences from common bread wheat. Because all the different classes of gliadins, α, β, γ, and ω, appear to be active in celiac disease, it is reasonably certain that the spelta gliadin is also toxic. We conclude that spelta is not a safe grain for people with celiac disease, contrary to the implications in labeling a bread made from spelta as “an alternative to wheat”. Our conclusions are in accord with spelta and bread wheat being classed taxonomically as subspecies of the same genus and species, Triticum aestivum L.  相似文献   

17.
《Cereal Chemistry》2017,94(3):546-553
Wheat proteins are classified according to solubility into the so‐called Osborne fractions. Because wheat flour contains both free thiol and disulfide groups, thiol–disulfide interchange reactions are possible during extraction. Osborne fractionation of 12 different wheat flour samples was performed in the presence of N‐ethylmaleinimide (NEMI) to alkylate free thiol groups and without addition of NEMI (control). The addition of NEMI during extraction tended to decrease the content of gliadins (predominantly α‐gliadins) and caused an increase of the content of glutenins in most flour samples. Thus, alkylation of free thiol groups during extraction led to a decline of the gliadin/glutenin ratio from 2 (control) to approximately 1.5 (NEMI). NEMI and control gliadins were separated by gel‐permeation HPLC into an oligomeric subfraction (high‐molecular‐weight [HMW] gliadins) and two monomeric subfractions. In most flours (8 of 12), the addition of NEMI led to a significant increase of the content of HMW gliadins. HMW gliadins from cultivar Akteur wheat were preparatively isolated from NEMI and control gliadins and characterized by HPLC, sodium dodecyl sulfate polyacrylamide gel electrophoresis, and N‐terminal sequencing. HMW gliadin isolated in the presence of NEMI had a significantly higher content of low‐molecular‐weight glutenin subunits and disulfide‐bound cysteine as well as a lower content of α‐gliadins and disulfide‐bound glutathione compared with the control.  相似文献   

18.
Use of saturated Ba(OH)2 to extract rye β‐glucan led to a depolymerized product. Similar depolymerization of β‐glucan was observed when oat bran was extracted with this reagent. Isolated oat β‐glucan, detarium xyloglucan, guar galactomannan, and wheat and rye arabinoxylan were also depolymerized by treatment with the barium reagent. The degree of depolymerization was related to time of contact with, and concentration of, the barium. Rye β‐glucan of two different molecular weights (MW) were isolated and characterized. The structure of rye β‐glucan, as evaluated from the ratio of (1→3)‐linked cellotriosyl to (1→3)‐linked cellotetraosyl primary structural units, most closely resembles barley β‐glucan. Analytical variability of this ratio is discussed. A freshly prepared solution (2%) of the higher MW sample showed shear thinning behavior typical of cereal β‐glucans. The lower MW sample at 2% was not shear thinning, but on further purification, after storage for seven days, a 6% solution had gelled as shown by the mechanical spectrum.  相似文献   

19.
诸葛菜OvCYP86MF基因的克隆及其特性分析   总被引:2,自引:1,他引:1  
为进一步阐明细胞色素P450基因CYP86MF在诸葛菜发育中的分子机理,本文根据已知同源基因保守序列设计特异引物,利用RT-PCR和RACE技术从诸葛菜中获得一个细胞色素P450基因(OvCYP86MF)全长cDNA序列,该序列全长为1876bp,含有1605bp的完整开放阅读框,可编码534个氨基酸,分子量和等电点分别为61.4kDa和6.90,具有细胞色素P450蛋白的典型特征,即保守结构域FNAGPRLCIG;原核表达显示该基因的融合蛋白在体外可以诱导表达;DANSTAR和Clustal W软件分析表明该基因的全长cDNA序列及其编码氨基酸序列与十字花科物种拟南芥相似性很高,达到80%以上,亲缘关系最近;与CYP86C亚家族成员在氨基酸水平上的相似性均高于50%,因此推断该基因属于CYP86C这个亚家族。Northern杂交分析表明该基因在花蕾中特异表达。  相似文献   

20.
The conventional Landry‐Moureaux method for selective extraction of maize proteins was modified by reducing the contact time of meal with extractants and by removing 55% 2‐propanol as extractant. The new procedure, coupled with a method for quantitating protein at microgram level, was used for assessing the nitrogen distribution of four soluble protein fractions present in 100‐mg samples of endosperm originating from six maize inbreds and opaque‐2 versions. Proteins extracted with 55% 2‐propanol plus reductant were made up of α‐, β‐, γ‐, and δ‐zeins. Proteins extracted subsequently with salt plus reductant were minor and poor in lysine (1 mol%).They were associated with zeins. Comparison of present data with those available in the literature showed a close similarity for a given genotype between the percentage of total α‐amino nitrogen extracted by 2‐propanol plus reductant than by salt plus reductant under conditions of the modified procedure and that of total Kjeldhal nitrogen extracted by 2‐propanol with and without reductant, and by salt plus reductant, using the conventional procedure. A simplified protocol was described and tested for isolating and quantitating α‐amino nitrogen as nonprotein, true protein, salt‐soluble proteins, zeins, and true glutelins in any sample of maize endosperm.  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