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 46 毫秒
1.
Cry1Ia and Cry1Aa proteins exhibited toxicities against Prays oleae with LC50 of 189 and 116 ng/cm2, respectively. The ability to process Cry1Ia11 protoxin by trypsin, chymotrypsin and P. oleae larvae proteases was studied and compared to that of Cry1Aa11. After solubilization under high alkaline condition (50 mM NaOH), Cry1Aa11 was converted into a major fragment of 65 kDa, whereas Cry1Ia11 protoxin was completely degraded by P. oleae larvae proteases and trypsin and converted into a major fragment of 70 kDa by chymotrypsin. Using less proteases of P. oleae juice, the degradation of Cry1Ia11 was attenuated. When the solubilization (in 50 mM Na2CO3 pH 10.5 buffer) and activation were combined, Cry1Ia11 was converted into a proteolytic product of 70 kDa after 3 h of incubation with trypsin, chymotrypsin and P. oleae juice. These results suggest that the in vivo solubilization of Cry1Ia11 was assured by larval proteases after a swelling of the corresponding inclusion due to the alkalinity of the larval midgut.  相似文献   

2.
Digestive proteinases and carbohydrases of Ectomyelois ceratoniae (Zeller) larvae were investigated using appropriate substrates and inhibitors. Midgut pH in larvae was determined to be slightly alkaline. Midgut extracts showed optimum activity for proteolysis of hemoglobin at pH 9–12. Midgut proteinases also hydrolyzed the synthetic substrates of trypsin, chymotrypsin, and elastase at pH 8–11. Maximum digestive α-amylase activity was also observed at pH 8–11. However, optimum activity for α- and β-glucosidase occurred at pH 5–8. Alpha- and β-galactosidases optimum activities occurred at pH 5 and pH 6, respectively. Inhibitors of serine proteases were effective on midgut serine proteases (trypsin and chymotrypsin proteases). Zymogram analyses revealed at least five bands of total proteolytic activity in the larval midgut. Protease-specific zymogram analyses revealed at least four, two, and one isozymes for trypsin-, chymotrypsin-, and elastase-like activities respectively. Two α-amylase isozymes were found in the midgut of fifth instar larvae and in the whole bodies of 1st through 5th instar larvae. Zymogram studies also revealed the presence of one and two bands of activity for β- and α-glucosidase, respectively. Recycling of α-amylase and proteases in the larval midgut was not complete. At least one isozyme of trypsin, chymotrypsin, elastase, and α-amylase were not recycled and were observed in the larval hindgut.  相似文献   

3.
A trypsin inhibitor purified from the seeds of the Manila tamarind, Pithecellobium dulce (PDTI), was studied for its effects on growth parameters and developmental stages of  Helicoverpa armigera. PDTI exhibited inhibitory activity against bovine trypsin (~86%; ~1.33 ug/ml IC50). The inhibitory activity of PDTI was unaltered over a wide range of temperature, pH, and in the presence of dithiothreitol. Larval midgut proteases were unable to digest PDTI for up to 12 h of incubation. Dixon and Lineweaver–Burk double reciprocal plots analysis revealed a competitive inhibition mechanism and a Ki of ~3.9 × 10?8 M. Lethal dose (0.50% w/w) and dosage for weight reduction by 50% (0.25% w/w) were determined. PDTI showed a dose‐dependent effect on mean larval weight and a series of nutritional disturbances. In artificial diet at 0.25% w/w PDTI, the efficiency of conversion of ingested food, of digested food, relative growth rate, and growth index declined, whereas approximate digestibility, relative consumption rate, metabolic cost, consumption index, and total developmental period were increased in larvae. This is the first report of antifeedant and antimetabolic activities of PDTI on midgut proteases of  H. armigera.  相似文献   

4.
We characterized trypsin‐ and chymotrypsin‐like serine alkaline proteases from cotton bollworm, Helicoverpa armigera, for their probable potential application as additives in various bio‐formulations. Purification was achieved by using hydroxylapatite, DEAE sephadex and CM sephadex columns, which resulted in increased enzyme activity by 13.76‐ and 14.05‐fold for trypsin and chymotrypsin, respectively. Michaelis–Menten constants (Km) for substrates of trypsin and chymotrypsin, BApNA and SAAPFpNA, were found to be 1.25 and 0.085 mM, correspondingly. Fluorescent zymogram analysis indicated the presence of five trypsin bands with molecular masses of ~21, 25, 38, 40, and 66 kDa and two chymotrypsin bands with molecular masses of ~29 and 34 kDa in SDS‐PAGE. The optimum pH was 10.0 and optimum temperature was 50°C for proteolytic activity for the purified proteases. The proteases were inhibited by synthetic inhibitors such as PMSF, aprotonin, leupeptin, pefabloc, and antipain. TLCK and TPCK inhibited about 94 and 90% of trypsin and chymotrypsin activity, respectively, while EDTA, EGTA, E64, pepstatin, idoacetamide, and bestatin did not affect the enzymes. The purified enzymes exhibited high stability and compatibility with metal ions; oxidizing, reducing, and bleaching agents; organic solvents; and commercial detergents. Short life cycles, voracious feeding behavior, and production of multiple forms of proteases in the midgut with rapid catalytic activity and chemostability can serve H. armigera as an excellent alternative source of industrially important proteases for use as additives in stain removers, detergents, and other bio‐formulations. Identification of enzymes with essential industrial properties from insect species could be a bioresource.  相似文献   

5.
Helicoverpa armigera is a polyphagous pest damaging vast numbers of different crops leading to decrease in total production. Use of Bt transgenic to control H. armigera has worked well but has increased resistance against Bt in H. armigera and controversies about the Bt transgenic making it imperative to find another strategy to control attack. Soybean is a nonhost plant for H. armigera; reason could be laid in the defense system of the soybean. Proteinase Inhibitor (PIs) have been extensively studied for development of resistance against insect pest. Two cultivars developed by our university were investigated for the presence of proteinase inhibitors namely, MAUS-158 and MAUS-61. Partially purified inhibitors were showed inhibition of total protease activity of gut extract by 91.34±1.49 and 89.95±0.96% by MAUS-158 and MAUS-61, respectively. While inhibition of trypsin like proteases were found between 65 and 71% and inhibition of chymotrypsin like proteases ranges between 40 and 42%. The partial purification study shows stability of PIs up to 60°C. Soybean PIs are also showing more prominent inhibition pattern against trypsin than chymotrypsin.  相似文献   

6.
The spectra of Tribolium castaneum and T. confusum larval digestive peptidases were characterized with respect to the spatial organization of protein digestion in the midgut. The pH of midgut contents in both species increased from 5.6–6.0 in the anterior to 7.0–7.5 in the posterior midgut. However, the pH optimum of the total proteolytic activity of the gut extract from either insect was pH 4.1. Approximately 80% of the total proteolytic activity was in the anterior and 20% in the posterior midgut of either insect when evaluated in buffers simulating the pH and reducing conditions characteristic for each midgut section. The general peptidase activity of gut extracts from either insect in pH 5.6 buffer was mostly due to cysteine peptidases. In the weakly alkaline conditions of the posterior midgut, the serine peptidase contribution was 31 and 41% in T. castaneum and T. confusum, respectively. A postelectrophoretic peptidase activity assay with gelatin also revealed the important contribution of cysteine peptidases in protein digestion in both Tribolium species. The use of a postelectrophoretic activity assay with p‐nitroanilide substrates and specific inhibitors revealed a set of cysteine and serine endopeptidases, 8 and 10 for T. castaneum, and 7 and 9 for T. confusum, respectively. Serine peptidases included trypsin‐, chymotrypsin‐, and elastase‐like enzymes, the latter being for the first time reported in Tenebrionid insects. These data support a complex system of protein digestion in the Tribolium midgut with the fundamental role of cysteine peptidases. © 2009 Wiley Periodicals, Inc.  相似文献   

7.
Insect midgut proteases are excellent targets for insecticidal agents such as protease inhibitors. These inhibitors are used for producing transgenic plants, resistant to pests. For achieving this goal, it is necessary to find the nature of specific proteases and their properties for adopting possible pest management procedure. Therefore, characterisation of the enzymes in the gut of the rose sawfly, Arge rosae (Hymenoptera: Argidae), responsible for proteolysis, was performed using a range of synthetic substrates and specific inhibitors. The optimum conditions for general proteases and trypsin were achieved at pH 10. The highest activity for general proteases was obtained at a temperature of 45°C. The use of specific inhibitors and SDS-PAGE (sodium dodecyl sulfate polyacrylamide gel electrophoresis) provided evidence to suggest that most of the proteases belonged to the serine group because of high inhibitory effect of phenyl methane sulfonyl fluoride on total proteolytic activity. Also, inhibition assays and zymogram analysis showed that metalloproteases are present in A. rosae digestive system. These results indicated that A. rosae larvae mainly used serine proteases for protein digestion, with chymotrypsin as the dominant form. The kinetic parameters of trypsin-like proteases using N-benzoyl-dl-arg-p-nitroanilide as substrate indicated that the K m and V max values of trypsin in the gut of the fifth instar larvae were 730 ± 17.3 μM and 456 ± 13.85 nmol min?1 mg?1 protein, respectively.  相似文献   

8.
We determined some biochemical properties of Oulema melanopus larval gut proteases. We found adult midgut enzyme preparations yielded results similar to whole‐larval preparations, permitting studies of the very small whole‐larval preparations. Protein preparations were analyzed using FITC–casein as a substrate. Acidic pH is optimal for proteolytic activity (range 3.0–4.0). Cysteine protease activity increased at acidic pH and in the presence of β‐mercaptoethanol. Protease activities at all pH values were maximal at 45°C. Enzyme activity in larval preparations was inhibited by addition of Fe2+, Ca2+, Mg2+, Zn2+, and K+ (10 mM). Fe2+ and Zn2+ significantly decreased enzyme activity at all pH values, Ca2+ and Mg2+ at pH 6.2 and Mg2+ at pH 4.0. Inhibitors, including pepstatin A, showed the greatest inhibition at pH 4.0; phenylmethylsulfonyl fluoride, N‐p‐tosyl‐l‐phenylalanine chloromethyl ketone at pH 6.2; and phenylmethylsulfonyl fluoride, Nα‐tosyl‐l‐lysine chloromethyl ketone hydrochloride, N‐p‐tosyl‐l‐phenylalanine chloromethyl ketone, trans‐epoxysuccinyl‐l‐leucylamido‐(4‐guanidino) butane at pH of 7.6. Inhibition assays indicated that cysteine, aspartyl (cathepsin D), serine (trypsin, chymotrypsin‐like) proteases and metalloproteases act in cereal leaf beetle digestion.  相似文献   

9.
Protease inhibitors play an important role in host plant defence against herbivores. However, insects have the ability to elevate the production of proteinases or resort to production of a diverse array of proteinases to offset the effect of proteinase inhibitors. Therefore, we studied the inhibition of pro‐proteinase(s) activation in the midgut of the polyphagous pest Helicoverpa armigera (Hübner) (Lepidoptera: Noctuidae) in response to protease inhibitors to develop appropriate strategies for the control of this pest. Gelatin coating present on X‐ray film was used as a substrate to detect electrophoretically separated pro‐proteinases and proteinases of H. armigera gut extract on native‐ and sodium dodecyl sulphate‐polyacrylamide gel electrophoresis. Six activated pro‐proteinase bands were detected in H. armigera gut lumen, which were partially purified and characterized using substrate assays. Activated H. armigera midgut pro‐proteinase(s) showed activity maxima at pH 8 and 10, and exhibited optimal activity at 40 °C. The activation of H. armigera gut pro‐proteinase isoforms was observed in the fraction eluted on benzamidine‐sepharose 4B column. Purification and substrate assay studies revealed that 23–70 kDa polypeptides were likely the trypsin/chymotrypsin‐like pro‐proteinases. Larvae of H. armigera fed on a cocktail of synthetic inhibitors (antipain, aprotinin, leupeptin, and pefabloc) showed maximum activation of pro‐proteinases compared with the larvae fed on individual inhibitors. The implications of these results for developing plants expressing proteinase inhibitors for conferring resistance to H. armigera are discussed.  相似文献   

10.
The major proteinase activities in the larval midgut of a common poplar tree borer, Anoplophora glabripennis, were characterised. Overall digestive capacity, as measured by casein hydrolysis, showed a pH optimum between 10 and 11.5, suggestive of serine endopeptidase activity. Trypsin, chymotrypsin, and chymotrypsin-like activities were detected using specific p-nitroanilide synthetic substrates and by use of specific serine endopeptidase inhibitors. These activities also showed pH optima in the extreme alkaline range. The absence of cysteine, aspartic, and metallo-endopeptidases were confirmed using class specific proteinase inhibitors. The dominant exopeptidase in the midgut is leucine aminopeptidase with a pH optimum of 7–9. Carboxypeptidase a and b activity were barely detectable. A large range of serine endopeptidase inhibitors were screened and were found to vary widely in their ability to inhibit casein hydrolysis. Potato proteinase inhibitor 1 (a chymotrypsin inhibitor) and wheat-germ trypsin inhibitor 1 inhibited particularly effectively in tandem and represent possible candidates for gene transformation to produce plants tolerant to this pest. © 1996 Wiley-Liss, Inc.  相似文献   

11.
Induced resistance in plants affects insect growth and development as a result of the up‐regulation of defence‐related secondary metabolites or enzyme‐binding proteins. In the present study, the effects of jasmonic acid (JA) and salicylic acid (SA) induced resistance in groundnut on Helicoverpa armigera (Hübner) are examined. Larval survival, larval weights and the activities of digestive enzymes (total serine protease and trypsin) and of detoxifying enzymes [glutathione S‐transferase (GST) and esterase (EST)] are studied in insects fed on four groundnut genotypes with moderate levels of resistance to H. armigera (ICGV 86699, ICGV 86031, ICG 2271 and ICG 1697) and a susceptible genotype (JL 24). The plants are pre‐ and/or simultaneously treated with JA and SA, and then infested with H. armigera, which are allowed to feed for 6 days. Significantly lower serine protease and trypsin activities are observed in H. armigera fed on plants treated with JA. Greater GST activity is recorded in insects fed on JA and SA treated plants, whereas EST activity is low in H. armigera larvae fed on plants treated with JA and SA. Serine proteases, trypsin and GST activities and larval weights (r = 0.74–0.95) and larval survival (r = 0.77–0.93) are positively correlated, whereas EST activity and larval weight (r = ?0.55) and larval survival (r = ?0.65) are negatively correlated. The results suggest that midgut digestive and detoxifying enzymes can be used as indicators of the adverse effects of constitutive and/or induced resistance in crop plants on the insect pests and the role of JA and SA in insect pest management.  相似文献   

12.
Chilo suppressalis is a key constraint on production of rice. The current research was conducted to study the types of digestive proteases in the larval midgut of C. suppressalis. It was found that activity of total digestive proteases increased from the first to the fifth larval instars, which showed different nutritional requirements. Four types of proteinases and two types of exopeptidase were identified so that their activities from the highest to the lowest activities is trypsin‐like, chymotrypsin‐like and elastase for proteinases, and amino and carboxypeptidases for exopeptidases. Meanwhile, just one type of cysteine protease, cathepsin D, was determined in the fourth and fifth instar larvae. The optimal pH for activity of total protease was found to be pH 9–10 and optimal temperature was observed to be 35–40°C, where there was the highest proteolytic activity. Some specific inhibitors of proteases including PMSF, TLCK, TPCK, DTT, E‐64, cystatin, phenanthroline and EDTA were used to confirm the types of proteases in the midgut of C. suppressalis.  相似文献   

13.
Abstract In the Ozark Mountains of the U.S.A., the red oak borer Enaphalodes rufulus contributes to the destruction of red oaks. To understand nutrient digestion in E. rufulus larvae, digestive proteinases are compared in both larvae fed heartwood phloem and those transferred to artificial diet. The pH of gut extracts is approximately 6.3 in the midgut and foregut and decreases to 5.5 in the hindgut region. The hydrolysis of casein by midgut extracts from E. rufulus larvae fed either artificial diet or phloem from tree sections increases in buffers greater than pH 6.19, with maximum hydrolysis being observed at pH 10.1. Casein zymogram analysis reveals two major proteinase activities in larval midgut extracts of diet‐fed larvae, with molecular masses of approximately 25 and 40–60 kDa, whereas phloem‐fed larvae have proteinase activities corresponding to 40, 45, 60, 80 and >100 kDa. Substrate analysis indicates at least one major trypsin‐like activity in both gut extracts with a molecular mass of >100 kDa, but two chymotrypsin‐like activities of approximately 25 and >200 kDa are found only in diet‐fed larvae. Inhibitors of serine proteinases are most effective in reducing the general proteolytic activity of midgut extracts from larvae fed either food source. The data indicate that serine proteinase inhibitors have the potential to reduce E. rufulus larval damage to oaks. In particular, transgenic technologies incoporating trypsin inhibitors may be effective in reducing protein digestion in phloem‐feeding larvae.  相似文献   

14.
Wai Leong  Joseph R. Pawlik 《Oikos》2010,119(9):1417-1422
Fragmentation and propagule formation are alternative reproductive strategies found in both plants and animals, with the latter generally providing greater dispersal capability. When both strategies occur, life history theory predicts that resources should be divided between the two. On coral reefs, both strategies are exhibited by branching corals and sponges, which are broken‐up after storm events and rapidly recolonize. In this study, we compared two congeneric Caribbean reef sponges, Callyspongia armigera, which is branched and easily fragmented, and C. vaginalis, which is not, to test whether there is a tradeoff in growth and propagule formation for C. armigera relative to C. vaginalis. Both species were equally abundant on coral reefs off Key Largo, Florida (10.1 ± 3.7 vs 11.9 ± 3.0 per 100 m2, respectively), suggesting that they are equally successful relative to two other non‐fragmenting congeneric species (C. fallax and C. plicifera) that are much less common. The number of substratum attachment points per sponge was significantly higher for C. armigera compared to C. vaginalis (2.31 ± 1.47 vs 1.03 ± 0.18 sponge?1), providing further evidence of the reliance of C. armigera on fragmentation, and of C. vaginalis on recruitment from larval settlement and subsequent growth. Growth rates in predator‐exclusion experiments were ~4‐fold higher for C. armigera compared to C. vaginalis (0.36 ± 0.31 vs 0.08 ± 0.11 % initial mass day?1), but C. armigera produced ~13‐fold fewer propagules than C. vaginalis (0.04 ± 0.22 vs 0.53 ± 1.08 % tissue area). Our results support a tradeoff between growth and propagule output for C. armigera relative to C. vaginalis, suggesting that these closely related sponge species took different evolutionary trajectories in reconciling their resource constraints.  相似文献   

15.
The adsorption, desorption and anti-ultraviolet light characteristics of the protoxin from Bacillus thuringiensis strain WG-001 on montmorillonite, kaolinite, zinc oxide and rectorite were studied. The protoxin was easily adsorbed onto minerals and the adsorption reached equilibrium within 0.5–1.0 h (except for rectorite). The adsorption isotherms of protoxin at different concentrations in sodium carbonate buffer (pH 9) followed the Langmuir (R 2 >0.97) and Freundlich (R 2 >0.95) equations. The maximum amounts of protoxin adsorbed were in the order: montmorillonite>rectorite>znic oxide>kaolinite. In the range of pH from 9 to 11 (carbonate buffer), the protoxin adsorbed decreased with increasing pH. The adsorption was not significantly affected by the temperature between 5 and 45°C. Both free and adsorbed protoxin were toxic to larvae of Heliothis armigera. The LC50 value of free and adsorbed protoxin on montmorillonite, rectorite, zinc oxide and kaolinite were 14±1.16, 1.76±0.31, 2.94±0.71, 4.78±2.08 and 1.91±0.91 µg mL?1, respectively. After 1 h of ultraviolet irradiation, the LC50 of the above samples increased by 41.4, 19.3, 16.3, 125.9 and 62.3%, respectively. The desorption of adsorbed protoxin in water ranged from 30.1 to 64.9% and from 18.5 to 48.7% in carbonate buffer.  相似文献   

16.
A trypsin inhibitor was purified from the seeds of Eugenia jambolana (Jambul) with a fold purification of 14.28 and a yield recovery of 2.8%. Electrophoretic analysis of E. jambolana trypsin inhibitor (EjTI) revealed a molecular weight of approximately 17.4 kDa on 12% denaturing polyacrylamide gel electrophoresis with or without reduction. EjTI exhibited high stability over a wide range of temperatures (4–80 °C for 30 min) and pH (3.0–10.0) and inhibited trypsin‐like activities of the midgut proteinases of fourth instar Helicoverpa armigera larvae by approximately 86%. Feeding assays containing 0.05, 0.15, and 0.45 (% w/w) EjTI on functionally important fourth‐instar larvae indicated a dose‐dependent downfall in the larval body weight as well as on extent of survival. The nutritional analysis suggests that EjTI exerts toxic effects on H. armigera. Dixon plot analysis revealed competitive inhibition of larval midgut proteinases by EjTI, with an inhibition constant (Ki) of approximately 3.1 × 10?9 M. However, inhibitor kinetics using double reciprocal plots for trypsin inhibition demonstrated a mixed inhibition pattern. These observations suggest the potential of E. jambolana trypsin inhibitor protein in insect pest management.  相似文献   

17.
棉铃虫4龄幼虫的摄食行为和中肠蛋白酶活性的变化   总被引:3,自引:0,他引:3  
李伟  王琛柱 《昆虫学报》1999,42(4):358-363
通过测定棉铃虫4龄幼虫,每隔12 h的摄食量、体重净增量以及丝氨酸蛋白酶活性的变化,揭示了它的取食行为及中肠蛋白酶活性的变化规律。结果表明:棉铃虫4龄幼虫取食及消化呈有规律的变化,蜕皮后4龄初,摄食量、体重净增量、类胰蛋白酶、类胰凝乳蛋白酶和总蛋白酶活性均逐渐增加;进入中期(36~48 h),各项指标达到最高值;其后,随着摄食量的下降,各项指标逐渐下降。同时, 连续观察幼虫取食高峰期的取食行为6 h,结果表明: 其间一次连续取食的平均时间为(3.6±0.4) min/次, 总平均取食时间为(62.4±8.9) min,约占整个时间的1/6。该文还对棉铃虫4龄幼虫不同时段取食行为与蛋白酶活性变化之间的相关性及调控机制作了初步讨论。  相似文献   

18.
Helicoverpa armigera is one of the most harmful pests in China. Although it had been successfully controlled by Cry1A toxins, some H. armigera populations are building up resistance to Cry1A toxins in the laboratory. Vip3A, secreted by Bacillus thuringiensis, is another potential toxin against H. armigera. Previous reports showed that activated Vip3A performs its function by inserting into the midgut brush border membrane vesicles (BBMV) of susceptible insects. To further investigate the binding of Vip3A to BBMV of H. armigera, the full-length Vip3Aa10 toxin expressed in Escherichia coli was digested by trypsin or midgut juice extract, respectively. Among the fragments of digested Vip3Aa10, only a 62 kDa fragment (Vip3Aa10-T) exhibited binding to BBMV of H. armigera and has insecticidal activity. Moreover, this interaction was specific and was not affected by the presence of Cry1Ab toxin. Binding of Vip3Aa10-T to BBMV resulted in the formation of an ion channel. Unlike Cry1A toxins, Vip3Aa10-T was just slightly associated with lipid rafts of BBMV. These data suggest that although activated Vip3Aa10 specifically interacts with BBMV of H. armigera and forms an ion channel, the mode of action of it may be different from that of Cry1A toxins.  相似文献   

19.
The distribution of digestive proteinases in either the anterior and posterior midgut or between the midgut epithelium and ectoperitrophic and endo-peritrophic spaces in the midgut were examined in the European corn borer, Ostrinia nubilalis. Trypsin, chymotrypsin, elastase, and aminopeptidase activities were the same in the anterior and posterior halves of the midgut. Of the total aminopeptidase activity, 95% was located in the midgut epithelium, and 90% of the trypsin, 97% of chymotrypsin, and 93% of the elastase activity were found in the midgut lumen. Trypsin, measured by hydrolysis of benzoyl-L-arginine ethyl ester, and chymotrypsin levels were significantly higher in the ectoperitrophic space compared to the endoperitrophic space. Digestion in the midgut is proposed to be sequential with tryptic digestion occurring in the endoperitrophic space. Ingested protein is digested further in the ectoperitrophic space by the action of elastase, chymotrypsin, and a second trypsin. Final digestion occurs by an intracellular aminopeptidase. © 1995 Wiley-Liss, Inc.  相似文献   

20.
The alfalfa weevil Hypera postica is a serious economic pest in most alfalfa grown in many countries worldwide. Digestive α-amylase and pectinase activities of larvae were investigated using general substrates. Midgut extracts from larvae showed an optimum activity for α-amylase against starch at acidic pH (pH 5.0). α-Amylase from larval midgut was more stable at mildly acidic pH (pH 5–6) than highly acidic and alkaline pH. The enzyme showed its maximum activity at 35°C. α-Amylase activity was significantly decreased in the presence of Ca2+, Mg2+ and sodium dodecylsulfate. On the contrary, K+ and Na+ did not significantly affect the enzyme activity. Zymogram analysis revealed the presence of one band of α-amylase activity in in-gel assays. Pectinase activity was assayed using agarose plate and colorimetric assays. Optimal pH for pectinase activity in the larval midgut was determined to be pH 5.0. Pectinase enzyme is more stable at pH 4.0–7.0 than highly acidic and alkaline pH. However, the enzyme was more stable at slightly acidic pH (pH 6.0) when incubation time increased. Maximum activity for the enzyme incubated at different temperatures was observed to be 40°C. Optimum pH activity for α-amylase and pectinase is not completely consistent with the pH prevailing in the larval midgut. This is the first report of the presence of pectinase activity in H. postica.  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