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.
AIMS: To analyse the cellular mechanisms that influence Listeria monocytogenes adhesion onto inert surfaces under acidic growth conditions. METHODS AND RESULTS: The adhesion capability of all the strains was significantly reduced after cultivation at constant pH 5 than at constant pH 7 and the cell surface was significantly less hydrophobic at pH 5 than at 7. At pH 5, the analyses of surface protein composition revealed that the flagellin was downregulated for all strains, which was confirmed by the absence of flagella and the P60 protein was upregulated for L. monocytogenes EGD-e, X-Li-mo 500 and 111. The use of L. monocytogenes EGD mutants revealed that flagellin could be involved in the adhesion process, but not P60 protein. It was also observed that the hydrophobic character was not linked to the presence or the absence of flagellin or P60 protein at the cell surface of L. monocytogenes. CONCLUSIONS: The decrease of L. monocytogenes adhesion at pH 5 could be attributed to the downregulation of the flagellin synthesis under the acidic conditions. SIGNIFICANCE AND IMPACT OF THE STUDY: Conservation of food product at pH 5 will delay bacterial adhesion and biofilm formation during food processing on inert surfaces when the product is contaminated with L. monocytogenes.  相似文献   

2.
Polarographic and UV-spectrophotometric investigations of Pb(II) complex formation with beta-cyclodextrin have showed that the complexation of Pb(II) ions begins at pH >10. The formation of lead(II) 1:1 complex with the beta-cyclodextrin anion was observed at pH 10-11.5. The logarithm of the stability constant of this complex compound is 15.9+/-0.3 (20 degrees C, ionic strength 1.0), and the molar extinction coefficient value is ca. 5500 (lambda(max)=260 nm). With further increase in solution pH the Pb-beta-cyclodextrin complex decomposes and converts to Pb(OH)(2) or Pb(OH)(3)(-) hydroxy-complexes. This process occurs with a decrease in Pb(II) complexation degree. The latter result could be explained by a decrease in the beta-cyclodextrin anion activity. Neither Pb(OH)(2) nor Pb(OH)(3)(-) encapsulation into beta-CD cavity was observed.  相似文献   

3.
Melanin is a highly irregular heteropolymer consisting of monomeric units derived from the enzymatic oxidation of the amino acid tyrosine. The process of melanin formation takes place in specialized acidic organelles (melanosomes) in melanocytes. The process of melanin polymerization requires an alkaline pH in vitro, and therefore, the purpose of an acidic environment in vivo remains a mystery. It is known that melanin is always bound to protein in vivo. It is also seen that polymerization in vitro at an acidic pH necessarily requires the presence of proteins. The effect of various model proteins on melanin synthesis and their interaction with melanin was studied. It was seen that many proteins could increase melanin synthesis at an acidic pH, and that different proteins resulted in the formation of different states of melanin, i.e., a precipitate or a soluble, protein-bound form. We also present evidence to show that soluble protein-bound melanin is present in vivo (in B16 cells as well as in B16 melanoma tissue). An acidic pH appeared to be necessary to ensure the formation of a uniform, very high molecular weight melano-protein complex. The interaction between melanin and proteins appears to be largely charge-dependent as evidenced by zeta potential measurements, and this interaction is also increased in an acidic pH. Thus, it appears that an acidic intramelanosomal pH is essential to ensure maximum interaction between protein and melanin, and also to ensure that all the melanin formed is protein-bound.  相似文献   

4.
Choi CY  Cerda JF  Chu HA  Babcock GT  Marletta MA 《Biochemistry》1999,38(51):16916-16924
Proteolysis of hemoglobin provides an essential nutrient source for the malaria parasite Plasmodium falciparum during the intraerythrocytic stage of the parasite's lifecycle. Detoxification of the liberated heme occurs through a unique heme polymerization pathway, leading to the formation of hemozoin. Heme polymerization has been demonstrated in the presence of P. falciparum histidine-rich protein 2 (PfHRP2) [Sullivan, D. J., Gluzman, I. Y., and Goldberg, D. E. (1996) Science 271, 219-221]; however, the molecular role that PfHRP2 plays in this polymerization is currently unknown. PfHRP2 is a 30 kDa protein composed of several His-His-Ala-His-His-Ala-Ala-Asp repeats and is present in the parasite food vacuole, the site of hemoglobin degradation and heme polymerization. We found that, at pH 7.0, PfHRP2 forms a saturable complex with heme, with a PfHRP2 to heme stoichiometry of 1:50. Spectroscopic characterization of heme binding by electronic absorption, resonance Raman, and EPR has shown that bound hemes share remarkably similar heme environments as >95% of all bound hemes are six-coordinate, low-spin, and bis-histidyl ligated. The PfHRP2-ferric heme complex at pH 5.5 (pH of the food vacuole) has the same heme spin state and coordination as observed at pH 7.0; however, polymerization occurs as heme saturation is approached. Therefore, formation of a PfHRP2-heme complex appears to be a requisite step in the formation of hemozoin.  相似文献   

5.
Formation of flavour compounds in the Maillard reaction   总被引:6,自引:0,他引:6  
This paper discusses the importance of the Maillard reaction for food quality and focuses on flavour compound formation. The most important classes of Maillard flavour compounds are indicated and it is shown where they are formed in the Maillard reaction. Some emphasis is given on the kinetics of formation of flavour compounds. It is concluded that the essential elements for predicting the formation of flavour compounds in the Maillard reaction are now established but much more work needs to be done on specific effects such as the amino acid type, the pH, water content and interactions in the food matrix. It is also concluded that most work is done on free amino acids but hardly anything on peptides and proteins, which could generate peptide- or protein-specific flavour compounds.  相似文献   

6.
The reaction of NADPH with the flavoenzyme mercuric reductase has been studied by rapid-scan stopped-flow spectrophotometry at 5 degrees C in the pH range 5.1-9.5. An intermediate formed within the dead time of the apparatus, and proposed to be an NADPH complex of oxidized enzyme, has an almost pH-independent spectrum. At pH 5.1 the formation of this species is followed by a rapid bleaching (k = 145 s-1) of the main flavin absorption band at 455 nm concomitantly with an absorbance increase around 395 nm. This process, which has a kinetic hydrogen isotope effect of 2.4, becomes less prominent at higher pH values and is not detectable above pH 7. It is suggested that this process includes the formation of a covalent thiol-flavin C-4a derivative stabilized by protonation of the active site. In the presence of an excess of NADPH, the final product of the reaction is probably an NADPH complex of two-electron-reduced enzyme, but below pH 6 the final spectrum becomes less intense suggesting a partial formation of four-electron-reduced enzyme. The spectral changes observed above pH 7 are nearly independent of pH. The first measurable step (k = 48 s-1 at pH 9.5) is thought to include the formation of an NADP+ complex of two-electron-reduced enzyme, while the final step (k = 6.3 s-1 at pH 9.5) results in the above-mentioned NADPH complex with two-electron-reduced enzyme. A minimal kinetic scheme rationalizing the observed pH dependence of the reaction and the observed isotope effects is presented.  相似文献   

7.
Extraction, concentration, and serological detection of staphylococcal enterotoxins from foods are laborious and time consuming. By exposing food extracts to an insoluble matrix tagged with specific anti-enterotoxin B, we have been able to recover the toxin from foods in a sensitive and rapid way. After mixing the reagents for 2 h at room temperature, immunoglobulin G antibodies were attached to CNBr-activated Sepharose 4B at pH 8.5 (0.1 M carbonate buffer with 0.5 M NaCl). Sepharose-antibody complex (1 ml) specifically recovered 0.1 to 30 mug of enterotoxin B from 400 ml of food extract (100 g of food) after mixing for 2 h at 4 C. The Sepharose-antibody-toxin complex was washed with 0.02 M phosphate-buffered saline at pH 7.2, and the toxin was dissociated by 2 to 4 ml of 0.2 M HCl-glycine plus 0.5 M NaCl buffer at pH 2.8. The recovered enterotoxin was free of interfering food components and could be detected serologically. Work to couple antibodies A, B, C, D, and E to Sepharose to recover all five toxins in one step is under study.  相似文献   

8.
Complexation equilibria between Al(III) and (−)-epigallocathechin gallate (EGCG) in the presence of acetate buffer have been studied by spectrophotometry. The method is based on the competition between EGCG and buffer ligands for Al(III) ions. The apparent formation constant of the EGCG complex for Al(III), which could be determined by measuring the absorbance of the free EGCG, decreased with increasing acetate ion concentration at a fixed pH. This phenomenon has been quantitatively investigated and both types of complexes (EGCG and acetate) could be analyzed. The apparent formation constant of Al(III) complex with EGCG also decreased with decreasing pH at a fixed acetate ion concentration. The pH dependence of the apparent formation constant indicates the 1:1 competition between metal ions and hydrogen ions for the binding site of EGCG. The intrinsic formation constant of Al-EGCG complex, the proton association constant of EGCG and the formation constant of Al-acetate complex are found to be log KAl-EGCG = 7.6 ± 0.1, log KH-EGCG = 7.65 ± 0.03 and log KAl-acetate = 2.07 ± 0.05 by a graphical analysis.  相似文献   

9.
The assembly of gelsolin with actin was followed by the increase of the fluorescence intensity of a fluorescence label bound to actin. The time course of the formation of the gelsolin-actin complex in the presence of micromolar [Ca2+] could be quantitatively interpreted by a model in which one actin molecule binds slowly to gelsolin in a rate-determining step and subsequently a second actin molecule is bound at least 40 times more rapidly. The rate of binding of the first actin molecule to gelsolin was found to be remarkably slow and to depend on the pH. The rate constants of formation of the gelsolin-actin complex range from 1.5 X 10(4) M-1 s-1 at pH 8 to 7 X 10(4) M-1 s-1 at pH 6.  相似文献   

10.
Copper containing cellulose material is of growing interest, e.g. offering alternative in the field of antimicrobials. Solutions of copper d-gluconate complexes (Cu(2+)-DGL) were used to introduce copper ions into a swollen cellulosic matrix. A ligand exchange mechanism forms the chemical basis of the sorption process. Copper sorption in cellulose was studied in the range between pH 6 and 13. An estimate for the complex stabilities of the Cu-cellulose system could be derived from the calculated species distribution of the different Cu(2+)-DGL complexes present. Spectrophotometry and cyclic voltammetry of Cu(2+)-DGL complex solution were used to confirm the presence of different species participating in the ligand exchange reaction. The pH dependent uptake of Cu(2+) ions in the cellulose matrix can be explained on the basis of the relative stabilities of Cu(2+)-DGL complex vs. Cu(2+)-cellulose complexes. In comparison to pH 10, higher copper content was observed at pH 6 and 13. Copper content was limited by carboxyl content of cellulosic materials, thus in analogy to the structure of Cu(2+)-DGL complexes participation of the carboxyl group as complex forming site is proposed. At high Cu(2+)-concentration and longer time of immersion in the copper complex solutions formation of solid deposits was observed on the surface of the treated fibres.  相似文献   

11.
The interaction between bovin β-Lactoglobulin (β-LG) and retinol at two different pH values was investigated by multispectroscopic, zeta potential, molecular modeling, and conductometry measurements. The steady state and polarization fluorescence spectroscopy revealed that complex formation at two different pH values could occur through a remarkable static quenching. According to fluorescence quenching, one set of binding site at pH 2 and two sets of binding sites at pH 7 were introduced for binding of retinol to β-LG that show the enhancement of saturation score of β-LG to retinol in dimmer condition. The polarization fluorescence analysis represented that there is more affinity between β-LG and retinol at pH 7 rather than at pH 2. The effect of retinol on β-LG was studied by UV-visible, circular dichroism (CD), and synchronous fluorescence, which indicated that retinol induced more structural changes on β-LG at pH 7. β-LG–retinol complex formation at two different pH values was recorded via applying resonance light scattering (RLS) and zeta potential. Conductometry and RLS showed two different behaviors of interaction between β-LG and retinol at two different pH values; therefore, dimmer formation played important roles in different behaviors of interaction between β-LG and retinol. The zeta potential was the implied combination of electrostatic and hydrophobic forces which are involved in β-LG–retinol complex at two different pH values, and the hydrophobic interactions play a dominant role in complex formation. Molecular modeling was approved by all experimental results. The acquired results suggested that monomer and dimmer states of β-LG can be induced by retinol with different behaviors.  相似文献   

12.
Functional group interactions involved in the formation of the glutamate dehydrogenase-NADPH binary complex have been studied by three independent but complementary approaches: the pH dependence of the overall dissociation constant measured by an improved differential spectroscopic technique; the pH dependence of the enthalpy of complex formation measured by flow calorimetry; and the pH dependence of the number of protons released to, or taken up from, the solvent in the complex formation reaction, measured by titration. We conclude that the coenzyme binds to the enzyme through three distinguishable interactions: a pH-independent process involving the binding of the reduced nicotinamide ring; a relatively weak "proton-stabilizing" process, occurring at low pH involving the shift at a pK of 6.3 in the free enzyme to 7.0 in the enzyme-NADPH complex; and a stronger "proton-destabilizing" process, occurring at a higher pH involving a shift of a pK of 8.5 in the enzyme down to 6.9 in the enzyme-NADPH complex. The proton ionization of the free enzyme involved in this third interaction exhibits some unusual thermodynamic parameters, having delta Go = +11.5 +/- 0.1 kcal mol-1, delta Ho = +19 +/- 1 kcal mol-1, and delta So = +23 eu. We show here that this proton ionization step is directly related to and indeed constitutes the "implicit" shift in enzyme macrostates which we have shown to be responsible for the existence of large highly nonlinear delta Cpo effects in the formation of this complex [Fisher, H. F., Colen, A. H., & Medary, R. T. (1981) Nature (London) 292, 271-272].  相似文献   

13.
Volatile emission during vertebrate decay is a complex process that is understood incompletely. It depends on many factors. The main factor is the metabolism of the microbial species present inside and on the vertebrate. In this review, we combine the results from studies on volatile organic compounds (VOCs) detected during this decay process and those on the biochemical formation of VOCs in order to improve our understanding of the decay process. Micro-organisms are the main producers of VOCs, which are by- or end-products of microbial metabolism. Many microbes are already present inside and on a vertebrate, and these can initiate microbial decay. In addition, micro-organisms from the environment colonize the cadaver. The composition of microbial communities is complex, and communities of different species interact with each other in succession. In comparison to the complexity of the decay process, the resulting volatile pattern does show some consistency. Therefore, the possibility of an existence of a time-dependent core volatile pattern, which could be used for applications in areas such as forensics or food science, is discussed. Possible microbial interactions that might alter the process of decay are highlighted.  相似文献   

14.
The antibiotic A23187 carries Ca2+ across Müller-Rudin membranes made from 1,2-dierucoyl-sn-glycero-3-phosphocholine and n-decane. The conductance of the membranes is not increased by the Ca2+-transport. The flux depends linearly on Ca2+ concentration and ionophore concentration (above pH 6). It increases with increasing pH, approximately by a factor of 4-5 between pH 6 and pH 8. Maximal Ca2+-fluxes of about 10(-10) mol-cm-2-s-1 were found. A counter transport of H+ could not be detected. The complex formation between A23187 and Ca2+ in egg phosphotidylcholine vesicles was studied spectroscopically. The results are consistent with the formation of a 2:1 complex. Optical absorption measurements on single phophatidylcholine membranes were used to calculate the concentration of membrane-bound ionophore A23187.  相似文献   

15.
Melanin is a highly irregular heteropolymer consisting of monomeric units derived from the enzymatic oxidation of the amino acid tyrosine. The process of melanin formation takes place in specialized acidic organelles (melanosomes) in melanocytes. The process of melanin polymerization requires an alkaline pH in vitro, and therefore, the purpose of an acidic environment in vivo remains a mystery. It is known that melanin is always bound to protein in vivo. It is also seen that polymerization in vitro at an acidic pH necessarily requires the presence of proteins. The effect of various model proteins on melanin synthesis and their interaction with melanin was studied. It was seen that many proteins could increase melanin synthesis at an acidic pH, and that different proteins resulted in the formation of different states of melanin, i.e., a precipitate or a soluble, protein‐bound form. We also present evidence to show that soluble protein‐bound melanin is present in vivo (in B16 cells as well as in B16 melanoma tissue). An acidic pH appeared to be necessary to ensure the formation of a uniform, very high molecular weight melano–protein complex. The interaction between melanin and proteins appears to be largely charge‐dependent as evidenced by zeta potential measurements, and this interaction is also increased in an acidic pH. Thus, it appears that an acidic intramelanosomal pH is essential to ensure maximum interaction between protein and melanin, and also to ensure that all the melanin formed is protein‐bound.  相似文献   

16.
Precipitation of enzymes with reversible soluble–insoluble polymers is a simple approach which can be easily scaled up. This work reports investigations aiming at verifying the existence of specific interactions and complex formation between porcine trypsin and poly acrylic acids using spectroscopy techniques. The trypsin–polymer complex was insoluble at pH lower than 5, with a stoichiometric ratio polymer mol per protein mol of 1:148. It took only a minute for the insoluble complex to form and it was redissolved modifying the pH of the medium. The enzymatic activity of trypsin was maintained even in the presence of the polymer and after precipitation poly acrylic acid presence protect the enzyme from itself degradation. The conditions of complex formation were studied using pure proteins that could be applied on porcine pancreas homogenates as an isolation strategy of trypsin.  相似文献   

17.
The effect of dextran sulfate on protein aggregation was investigated to provide the clues of its biochemical mechanism. The interaction between dextran sulfate and BSA varied with the pH values of the solution, which led to the different extent of aggregation prevention by dextran sulfate. Light scattering data with thermal scan showed that dextran sulfate suppressed BSA aggregation at pH 5.1 and pH 6.2, while it had no effect at pH 7.5. Isothermal titration calorimetric analysis suggested that the pH dependency of the role of dextran sulfate on BSA aggregation would be related to the difference in the mode of BSA-dextran sulfate complex formation. Isothermal titration calorimetric analysis at pH 6.2 indicated that dextran sulfate did not bind to native BSA at this pH, but interacted with partially unfolded BSA. While stabilizing native form of protein by the complex formation has been suggested as the suitable mechanism of preventing aggregation, our observation of conformational changes by circular dichroism spectroscopy showed that strong electrostatic interaction between dextran sulfate and BSA rather facilitated the denaturation of BSA. Combining the data from isothermal titration calorimetry, circular dichroism, and dynamic light scattering, we found that the complex formation of the intermediate state of denatured BSA with dextran sulfate is a prerequisite to suppress the aggregation by preventing further oligomerization/aggregation process of denatured protein.  相似文献   

18.
Reduced azurin reacts with the resting, oxidized cytochrome c peroxidase of Pseudomonas aeruginosa to yield time courses observed at 420 nm, which consist of the sum of two exponential processes. Each process exhibits a hyperbolic dependence of the observed rate constant on the reduced azurin concentration. The fraction of the total optical density change which each process contributes is found to be dependent on the reduced azurin concentration. This pattern of reactivity is maintained at pH values between 5.5 and 8.0. The data has been analyzed in terms of a complex formation between the two proteins followed by an intramolecular electron exchange reaction. This analysis yields values for the binding constants at each pH value. The intramolecular exchange reaction is independent of pH, whilst the pH dependence of the binding reaction suggests the involvement of a histidine residue in this process.  相似文献   

19.
Dental microwear formation on the posterior dentition is largely attributed to an organism's diet. However, some have suggested that dietary and environmental abrasives contribute more to the formation process than food, calling into question the applicability of dental microwear to the reconstruction of diet in the fossil record. Creating microwear under controlled conditions would benefit this debate, but requires accurately replicating the oral environment. This study tests the applicability of Artificial Resynthesis Technology (ART 5) to create microwear textures while mitigating the challenges of past research. ART 5 is a simulator that replicates the chewing cycle, responds to changes in food texture, and simulates the actions of the oral cavity. Surgically extracted, occluding pairs of third molars (n = 2 pairs) were used in two chewing experiments: one with dried beef and another with sand added to the dried beef. High-resolution molds were taken at 0, 50, 100, 2500, and 5000 simulated chewing cycles, which equates to approximately 1 week of chewing. Preliminary results show that ART 5 produces microwear textures. Meat alone may produce enamel prism rod exposure at 5000 cycles, although attrition cannot be ruled out. Meat with sand accelerates the wear formation process, with enamel prism rods quickly obliterated and “pit-and-scratch” microwear forming at approximately 2500 cycles. Future work with ART 5 will incorporate a more thorough experimental protocol with improved controls, pH of the simulated oral environment, and grit measurements; however, these results indicate the potential of ART 5 in untangling the complex variables of dental microwear formation.  相似文献   

20.
1. The reconstitution of chlorophyllide biosynthesis by barley etioplast membranes is described. 2. The process is dependent on the additon of NADPH and protochlorophyllide and on illumination, which can be either continuous or intermittent. 3. The reconstituted process involves spectroscopically similar intermediates to the native reaction in whole leaves. 4. Steps in the process are an initial enzymic formation in the dark of a photoactive complex, P638/652 (probably a ternary protochlorophyllide-NADPH-enzyme complex), followed by a very rapid light-dependent hydrogen transfer from the NADPH to the protochlorophyllide giving chlorophyllide giving chlorophyllide, finally releasing the enzyme for repeating the process. 5. A continuous assay for the system regenerating complex P638/652 was devised on the basis of monitoring chlorophyllide formation. 6. The pH optimum of the reaction is at 6.9 and Km values for protochlorophyllide and NADPH are 0.46 and 35 micron respectively. 7. The reaction is associated specifically with the etioplast membrane fraction. 8. Activities of the system assayed in vitro are more than adequate to account for rates of chlorophyll formation in vivo.  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