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 15 毫秒
1.
The atomic force microscope (AFM) and the transmission electron microscope (TEM) have been used to study the morphology of isolated mouse thymocyte microdomains and Thy-1 antigen distribution at the surface of these structures. AFM images were recorded in air in the contact mode on membrane vesicles deposited on previously heated tissue culture plastic sheets and indirectly immunolabeled for Thy-1 expression with colloidal gold-conjugated secondary antibodies. AFM images of untreated plastic plates showed a very characteristic network of streaks 20-200 nm wide. Heating the plastic removed the streaks and provided flat surfaces (r.m.s. 1 nm). This substrate allowed strong adsorption and homogeneous spreading of the vesicles and easy manipulations during immunolabeling experiments. Vesicles flattened on the substrate without losing their morphology. The 10-nm membrane-bound gold beads were reproducibly imaged without degradation by repeated tip scanning. The observed microdomains had a mean diameter of 184 +/- 76 nm, and 65% of them were specifically labeled. Images obtained with the TEM on the same vesicles, deposited on carbon-coated grids and negatively stained, confirmed the AFM observations. The size distribution of the microdomains was quite similar, but the number of beads per vesicle was significantly higher, and 76% of the vesicles were labeled. The difference may be explained 1) by removal of beads from the vesicles in the additional washing step with water, which was necessary for the AFM; 2) by tip-sample convolution; and 3) by statistical fluctuations.  相似文献   

2.
Morin-aluminium fluorescence and membrane filtration were successfully applied to the indirect identification of the formation of hydroxyaluminosilicates (HAS) in acidic solutions of varying pH and of known concentrations of aluminium (Al) and silicic acid (Si(OH)(4)). It was proven to be especially useful in providing evidence of the strong competition between Si(OH)(4) and Al(OH)(3) to condense with hydroxyaluminium templates to form HAS in preference to Al(OH)(3(s)). The aggregation and stability of HAS and Al(OH)(3(s)) were dependent upon both the pH and the [Al] of the solution. The applicability of these indirect techniques was confirmed using the direct observation of HAS in solution by atomic force microscopy (AFM). AFM was also a powerful tool in providing valuable information on the morphology of colloidal HAS of various structures and stoichiometries. The results have provided further confirmation of both the mechanism of HAS formation and the form and stability of HAS in solution. This information is essential to our understanding of the biological availability and hence toxicity of Al in biota, including man.  相似文献   

3.
We tested the ability of saturated n-monocarboxylic acids ranging from eight to 12 carbons in length to self-assemble into vesicles, and determined the minimal concentrations and chain lengths necessary to form stable bilayer membranes. Under defined conditions of pH and concentrations exceeding 150 mM, an unbranched monocarboxylic acid as short as eight carbons in length (n-octanoic acid) assembled into vesicular structures. Nonanoic acid (85 mM) formed stable vesicles at pH 7.0, the pK of the acid in bilayers, and was chosen for further testing. At pH 6 and below, the vesicles were unstable and the acid was present as droplets. At pH ranges of 8 and above clear solutions of micelles formed. However, addition of small amounts of an alcohol (nonanol) markedly stabilized the bilayers, and vesicles were present at significantly lower concentrations (approximately 20 mM) at pH ranges up to 11. The formation of vesicles near the pK(a) of the acids can be explained by the formation of stable RCOO(-)...HOOCR hydrogen bond networks in the presence of both ionized and neutral acid functions. Similarly, the effects of alcohols at high pH suggests the formation of stable RCOO(-)...HOR hydrogen bond networks when neutral RCOOH groups are absent. The vesicles provided a selective permeability barrier, as indicated by osmotic activity and ionic dye capture, and could encapsulate macromolecules such as DNA and a protein. When catalase was encapsulated in vesicles of decanoic acid and decanol, the enzyme was protected from degradation by protease, and could act as a catalyst for its substrate, hydrogen peroxide, which readily diffused across the membrane. We conclude that membranous vesicles produced by mixed short chain monocarboxylic acids and alcohols are useful models for testing the limits of stabilizing hydrophobic effects in membranes and for prebiotic membrane formation.  相似文献   

4.
The yields of nitrate and nitrite from decomposition of peroxynitrite in phosphate buffer at 37 degrees C were determined in the pH range 1-14. The NO(2)(-)/NO(3)(-) yields showed a stepwise variation with pH, with inflection points at approximately pH 3.1, 5.8, 6.8, 8.0, and 11.9. Nitrite formation increased strongly above pH 7 at the expense of nitrate, but above pH 12 nitrate again became the major product (80% at pH 14). At this pH, the Arrhenius parameters were E(a)=24.1+/-0.2kcal mol(-1) and A=(4.9+/-1.3)x10(12)s(-1). The yields of NO(2)(-), NO(3)(-), and O(2) measured at pH 5.8, 7.4, and 8.5 as a function of the initial peroxynitrite concentration (50-1000 microM) were linear only at pH 5.8. In the presence of carbon dioxide, oxygen production at pH 7.5 and pH 10 was found to be linear on the CO(2) concentration. The experimental observations were satisfactorily reproduced by kinetic simulations including principal component analyses. These data strongly suggest that the chemistry of peroxynitrite is exclusively mediated by z.rad;NO(2) and HO(z.rad;) radicals in the absence, and by z.rad;NO(2) and CO(3)(z.rad;-) radicals in the presence of CO(2).  相似文献   

5.
Conventional transmission electron microscopy (TEM) was utilized to characterize vesicles formed by the spontaneous self-assembly of a novel zwitterionic block copolymer in the ionic liquid (2-hydroxyethyl)dimethylammonium methanesulfonate as well as in 0.1 M phosphate buffered saline (PBS). This block copolymer was synthesized via ring-opening metathesis polymerization (ROMP) of a norbornene-based sulfobetaine, followed by its end-functionalization with polystyrene to generate the necessary amphiphilic structure. The ionic liquid enabled the visualization of the vesicles in their swollen state by TEM, demonstrating a new method for improved characterization of polymer vesicles.  相似文献   

6.
We tested the ability of saturated n-monocarboxylic acids ranging from eight to 12 carbons in length to self-assemble into vesicles, and determined the minimal concentrations and chain lengths necessary to form stable bilayer membranes. Under defined conditions of pH and concentrations exceeding 150 mM, an unbranched monocarboxylic acid as short as eight carbons in length (n-octanoic acid) assembled into vesicular structures. Nonanoic acid (85 mM) formed stable vesicles at pH 7.0, the pK of the acid in bilayers, and was chosen for further testing. At pH 6 and below, the vesicles were unstable and the acid was present as droplets. At pH ranges of 8 and above clear solutions of micelles formed. However, addition of small amounts of an alcohol (nonanol) markedly stabilized the bilayers, and vesicles were present at significantly lower concentrations (∼20 mM) at pH ranges up to 11. The formation of vesicles near the pKa of the acids can be explained by the formation of stable RCOO…HOOCR hydrogen bond networks in the presence of both ionized and neutral acid functions. Similarly, the effects of alcohols at high pH suggests the formation of stable RCOO…HOR hydrogen bond networks when neutral RCOOH groups are absent. The vesicles provided a selective permeability barrier, as indicated by osmotic activity and ionic dye capture, and could encapsulate macromolecules such as DNA and a protein. When catalase was encapsulated in vesicles of decanoic acid and decanol, the enzyme was protected from degradation by protease, and could act as a catalyst for its substrate, hydrogen peroxide, which readily diffused across the membrane. We conclude that membranous vesicles produced by mixed short chain monocarboxylic acids and alcohols are useful models for testing the limits of stabilizing hydrophobic effects in membranes and for prebiotic membrane formation.  相似文献   

7.
The kinetics of the oxidation of D-glucose to D-gluconic acid by bromine in aqueous solution were studied using potentiometric techniques and theoretical considerations of complex bromine-bromide-pH equilibria. The pH has a strong influence on reaction rate. At pH < 8 the reaction is very slow, while in the pH range pH 8-9.5 the reaction is sufficiently fast and seems optimal for the reaction. The proposed active species at that pH region is hypobromous acid. At pH > 9.5, the reaction is further accelerated due to the formation of hypobromite. The proposed kinetics expression for gluconic acid formation, based on the determined kinetic parameters at pH 9.24, is of the form dc(GA)/dt = 160c(2)(G)c(o)(HOBr)c(o)(H(+)c(o)(Br)  相似文献   

8.
Chin TM  Lin SB  Lee SY  Chang ML  Cheng AY  Chang FC  Pasternack L  Huang DH  Kan LS 《Biochemistry》2000,39(40):12457-12464
The formation of a DNA "paper-clip" type triple helix (triplex) with a common sequence 5'-d-(TC)(3)T(a)()(CT)(3)C(b)()(AG)(3) (a and b = 0-4) was studied by UV thermal melting experiments and CD spectra. These DNA oligomers form triplexes and duplexes under slightly acidic and neutral conditions, respectively. The stability of the formed triplexes (at pH 4.5) or duplexes (at pH 7.0 or 8.0) does not vary significantly with the size of the loops (a and b = 1-4). At pH 6.0, the triplex stability is, however, a function of a and b. It is also interesting to note that the oligomer 5'-d-(TC)(3)(CT)(3)(AG)(3) (a and b = 0) forms a stable triplex at pH 4.5 with a slightly lower T(m) value, due to dissociation of a base triad at one end and a distorted base triad at the other, observed by (1)H NMR. Thus, we have here a model system, 5'-d-(TC)(3)T(a)(CT)(3)C(b)(AG)(3), that could form a triplex effectively with (a and b = 1-4) and without (a and b = 0) loops under acidic conditions. In addition, the triplex formation of oligomers with replacement of one, two, or three 2'-deoxycytidine in the Hoogsteen strand by either 2'-deoxypseudoisocytidine (D) or 2'-O-methylpseudoisocytidine (M) was also studied in the sequence 5'-d-(TX)(3)T(2)(CT)(3)C(2)(AG)(3) (where X is C, D, or M). Both CD spectra and UV melting results showed that only D3 [(TX)(3) = (TD)(3)] and M3 [(TX)(3) = (TM)(3)] were able to form the paper-clip structure under both neutral and acidic conditions. This is because the N(3)H of a pseudoisocytosine base can serve as a proton donor without protonation. We hereby proved that the 2'-deoxypseudoisocytidine, similar to 2'-O-methylpseudoisocytidine, could replace 2'-deoxycytidine in the Hoogsteen strand to provide triplex formation at neutral pH.  相似文献   

9.
The aim of the present study was to design amphiphilic oligopeptides that can self-assemble into vesicular structures. The ratio of hydrophilic to hydrophobic block length was varied, and peptides were designed to have a hydrophobic tail in which the bulkiness of the amino acid side groups increases toward the hydrophilic domain (Ac-Ala-Ala-Val-Val-Leu-Leu-Leu-Trp-Glu(2/7)-COOH). These peptides were recombinantly produced in bacteria as an alternative to solid-phase synthesis. We demonstrate with different complementary techniques (dynamic and static light scattering, tryptophan fluorescence anisotropy, and electron microscopy) that these amphiphilic peptides spontaneously form vesicles with a radius of approximately 60 nm and a low polydispersity when dispersed in aqueous solution at neutral pH. Morphology and size of the vesicles were relatively insensitive to the variations in hydrophilic block length. Exposure to acidic pH resulted in formation of visible aggregates, which could be fully reversed to vesicles upon pH neutralization. In addition, it was demonstrated that water-soluble molecules can be entrapped inside these peptide vesicles. Such peptide vesicles may find applications as biodegradable drug delivery systems with a pH-dependent release profile.  相似文献   

10.
Matrix vesicles are extracellular organelles involved in mineral formation that are regulated by 1alpha,25(OH)(2)D(3). Prior studies have shown that protein kinase C (PKC) activity is involved in mediating the effects of 1alpha,25(OH)(2)D(3) in both matrix vesicles and plasma membranes. Here, we examined the regulation of matrix vesicle PKC by 1alpha,25(OH)(2)D(3) during biogenesis and after deposition in the matrix. When growth zone costochondral chondrocytes were treated for 9 min with 1alpha,25(OH)(2)D(3), PKCzeta in matrix vesicles was inhibited, while PKCalpha in plasma membranes was increased. In contrast, after treatment for 12 or 24 h, PKCzeta in matrix vesicles was increased, while PKCalpha in plasma membranes was unchanged. The effect of 1alpha,25(OH)(2)D(3) was stereospecific and metabolite-specific. Monensin blocked the increase in matrix vesicle PKC after 24 h, suggesting the secosteroid-regulated packaging of PKC. In addition, the 1alpha,25(OH)(2)D(3) membrane vitamin D receptor (1,25-mVDR) was involved, since a specific antibody blocked the 1alpha,25(OH)(2)D(3)-dependent changes in PKC after both long and short treatment times. In contrast, antibodies to annexin II had no effect, and there was no evidence for the presence of the nuclear VDR on Western blots. To investigate the signaling pathways involved in regulating matrix vesicle PKC activity after biosynthesis, matrix vesicles were isolated and then treated for 9 min with 1alpha,25(OH)(2)D(3) in the presence and absence of specific inhibitors. Inhibition of phosphatidylinositol-phospholipase C, phospholipase D, or G(i)/G(s) had no effect. However, inhibition of G(q) blocked the effect of 1alpha,25(OH)(2)D(3). The rapid effect of 1alpha,25(OH)(2)D(3) also involved the 1,25-mVDR. Moreover, arachidonic acid was found to stimulate PKC when added directly to isolated matrix vesicles. These results indicate that matrix vesicle PKC is regulated by 1alpha,25(OH)(2)D(3) at three levels: 1) during matrix vesicle biogenesis; 2) through direct action on the membrane; and 3) through production of other factors such as arachidonic acid.  相似文献   

11.
Inorganic arsenic is an established human carcinogen, but its metabolism is incompletely defined. The ATP binding cassette protein, multidrug resistance protein (MRP1/ABCC1), transports conjugated organic anions (e.g. leukotriene C(4)) and also co-transports certain unmodified xenobiotics (e.g. vincristine) with glutathione (GSH). MRP1 also confers resistance to arsenic in association with GSH; however, the mechanism and the species of arsenic transported are unknown. Using membrane vesicles prepared from the MRP1-overexpressing lung cancer cell line, H69AR, we found that MRP1 transports arsenite (As(III)) only in the presence of GSH but does not transport arsenate (As(V)) (with or without GSH). The non-reducing GSH analogs L-gamma-glutamyl-L-alpha-aminobutyryl glycine and S-methyl GSH did not support As(III) transport, indicating that the free thiol group of GSH is required. GSH-dependent transport of As(III) was 2-fold higher at pH 6.5-7 than at a more basic pH, consistent with the formation and transport of the acid-stable arsenic triglutathione (As(GS)(3)). Immunoblot analysis of H69AR vesicles revealed the unexpected membrane association of GSH S-transferase P1-1 (GSTP1-1). Membrane vesicles from an MRP1-transfected HeLa cell line lacking membrane-associated GSTP1-1 did not transport As(III) even in the presence of GSH but did transport synthetic As(GS)(3). The addition of exogenous GSTP1-1 to HeLa-MRP1 vesicles resulted in GSH-dependent As(III) transport. The apparent K(m) of As(GS)(3) for MRP1 was 0.32 microM, suggesting a remarkably high relative affinity. As(GS)(3) transport by MRP1 was osmotically sensitive and was inhibited by several conjugated organic anions (MRP1 substrates) as well as the metalloid antimonite (K(i) 2.8 microM). As(GS)(3) transport experiments using MRP1 mutants with substrate specificities differing from wild-type MRP1 suggested a commonality in the substrate binding pockets of As(GS)(3) and leukotriene C(4). Finally, human MRP2 also transported As(GS)(3). In conclusion, MRP1 transports inorganic arsenic as a tri-GSH conjugate, and GSTP1-1 may have a synergistic role in this process.  相似文献   

12.
Fusion pores or porosomes are basket-like structures at the cell plasma membrane, at the base of which, membrane-bound secretory vesicles dock and fuse to release vesicular contents. Earlier studies using atomic force microscopy (AFM) demonstrated the presence of fusion pores at the cell plasma membrane in a number of live secretory cells, revealing their morphology and dynamics at nm resolution and in real time. ImmunoAFM studies demonstrated the release of vesicular contents through the pores. Transmission electron microscopy (TEM) further confirmed the presence of fusion pores, and immunoAFM, and immunochemical studies demonstrated t-SNAREs to localize at the base of the fusion pore. In the present study, the morphology, function, and composition of the immunoisolated fusion pore was investigated. TEM studies reveal in further detail the structure of the fusion pore. Immunoblot analysis of the immunoisolated fusion pore reveals the presence of several isoforms of the proteins, identified earlier in addition to the association of chloride channels. TEM and AFM micrographs of the immunoisolated fusion pore complex were superimposable, revealing its detail structure. Fusion pore reconstituted into liposomes and examined by TEM, revealed a cup-shaped basket-like morphology, and were functional, as demonstrated by their ability to fuse with isolated secretory vesicles.  相似文献   

13.
Surfactin is a bacterial lipopeptide with powerful surfactant-like properties. High-sensitivity isothermal titration calorimetry was used to study the self association and membrane partitioning of surfactin. The critical micellar concentration (CMC), was 7.5 microM, the heat of micellization was endothermic with DeltaH(w-->m)(Su) = +4.0 kcal/mol, and the free energy of micellization DeltaG(O,w-->m)(Su) = -9.3 kcal/mol (25 degrees C; 100 mM NaCl; 10 mM TRIS, 1 mM EDTA; pH 8.5). The specific heat capacity of micellization was deduced from temperature dependence of DeltaH(w-->m)(Su) as DeltaC(w-->m)(P) = -250 +/- 10 cal/(mol.K). The data can be explained by combining the hydrophobicity of the fatty acyl chain with that of the hydrophobic amino acids. The membrane partition equilibrium was studied using small (30 nm) and large (100 nm) unilamellar POPC vesicles. At 25 degrees C, the partition coefficient, K, was (2.2 +/- 0.2) x 10(4) M(-1) for large vesicles leading to a free energy of DeltaG(O, w-->b)(Su) = -8.3 kcal/mol. The partition enthalpy was again endothermic, with DeltaH(w-->b)(Su) = 9 +/- 1 kcal/mol. The strong preference of surfactin for micelle formation over membrane insertion explains the high membrane-destabilizing activity of the peptide. For surfactin and a variety of non-ionic detergents, the surfactant-to-lipid ratio, inducing membrane solubilization, R(sat)(b), can be predicted by the simple relationship R(sat)(b) approximately K. CMC.  相似文献   

14.
Rat tail tendon (RTT) collagen has been reacted with a homologous series of chromium(III) complexes viz., (H2O)(4)Cr(OH)(2)Cr(H2O)(4+)(4) 1 (dimer), Cr(3)(OH)(4)(H2O)(5+)(9) 2 (trimer), and Cr(4)(OH)(4)(O2)(H2O)(4+)(12) 3 (tetramer), and the structural alterations brought about by these complexes have been investigated through atomic force microscopy (AFM) and circular dichroism (CD) studies. Examination of Cr(III)-treated tendons using AFM revealed changes in the D-periodicity of collagen, which may arise due to differences in the topological distribution of various Cr(III) complexes. Evidence for organisation of monomeric collagen into quarter staggered fibrils in the presence of Cr(III) dimer, 1, has been obtained. The quaternary structural changes induced by chromium in the protein have been correlated to the conformational changes of collagen in the absence of denaturation.  相似文献   

15.
AFM of biological material embedded in epoxy resin   总被引:1,自引:0,他引:1  
We present a simple method to extract morphological details from the block face of epoxy embedded biopolymers by AFM. It is shown that topographical contrast and the identification of small structural details critically depend on the procedure of sample preparation before embedding (chemical fixation or high-pressure freezing and freeze-substitution) and on the hardness of the embedding epoxy resin. Ethanol treatment of the block face of the sample after microtomy elutes non-cross-linked polymer chains and makes the smallest details of the embedded biomaterial amenable to detection. AFM (height and phase contrast) examination of the block face of accordingly prepared cells of Caenorhabditis elegans provides data that are comparable to TEM.  相似文献   

16.
Bothrojaracin (BJC) is a 27-kD snake venom protein from Bothrops jararaca that has been characterized as a potent thrombin inhibitor. BJC binds to exosites I and II, with a dissociation constant of 0.7 nM, and influences but does not block the proteinase catalytic site. BJC also binds prothrombin through an interaction that has not been characterized. In the present work we characterize the interaction of BJC with prothrombin quantitatively for the first time, and identify the BJC binding site on human prothrombin. Gel filtration chromatography demonstrated calcium-independent, 1:1 complex formation between fluorescein-labeled BJC ([5F]BJC) and prothrombin, whereas no interactions were observed with activation fragments 1 or 2 of prothrombin. Isothermal titration calorimetry showed that binding of BJC to prothrombin is endothermic, with a dissociation constant of 76 +/- 32 nM. The exosite I-specific ligand, hirudin(54-65) (Hir(54-65) (SO(3)(-)), displaced competitively [5F]BJC from prothrombin. Titration of the fluorescent hirudin(54-65) derivative, [5F]Hir(54-65)(SO(3)(-)), with human prothrombin showed a dissociation constant of 7.0 +/- 0.2 microM, indicating a approximately 100-fold lower binding affinity than that exhibited by BJC. Both ligands, however, displayed a similar, approximately 100-fold increase in affinity for exosite I when prothrombin was activated to thrombin. BJC efficiently displaced [5F]Hir(54-65)(SO(3)(-)) from complexes formed with thrombin or prothrombin with dissociation constants of 0.7 +/- 0.9 nM and 11 +/- 80 nM, respectively, indicating that BJC and Hir(54-65)(SO(3)(-)) compete for the same exosite on these molecules. The results indicate that BJC is a potent and specific probe of the partially exposed anion-binding exosite (proexosite I) of human prothrombin.  相似文献   

17.
Secondary structure formation in four novel hybrid poly(acrylic acid)-b-poly(L-valine) (PAA-b-PLVAL) block copolymers, that is, PAA(40)-PLVAL(100), PAA(80)-PLVAL(100), PAA(80)-PLVAL(80), and PAA(80)-PLVAL(60), was investigated by circular dichroism. The formation of stable and well-defined beta-sheet structure in the PLVAL hydrophobic domains was observed for all the copolymers. At pH 5, PAA(80)-PLVAL(60) with the lowest PLVAL/PAA molar ratio possessed the lowest beta-sheet content of 12%, and it increased to 62% for PAA(40)-PLVAL(100) system. The beta-sheet formation in the block copolymers was controlled by both random PAA-PLVAL hydrogen bonds at low pH and electrostatic repulsive forces on the PAA segment at high pH; hence, the beta-sheet structure was most stable at intermediate pH. The length of PAA segments was critical in the beta-sheet solubilization and in providing sufficient shielding of the hydrophobic core from denaturing agents such as urea.  相似文献   

18.
The self-assembling properties of a series of single-chain (C12–C18) amphiphilic ligands, diazafluorenone Schiff bases (DAFSB), were studied in dilute aqueous solutions by various physical methods. Transmission electron microscopy (TEM) shows that these amphiphiles can form vesicles with diameters of 50–250 nm and layer widths of about 5 nm. UV-vis spectra reflect the formation of J-like aggregates in bilayer assemblies. The gel to liquid-crystal phase-transition behavior of the bilayer in vesicles was investigated by differential scanning calorimetry (DSC), and the phase transition temperature,T m, ranged between 60 and 75 °C. The experimental results indicate that DAFSB is a new type of bilayer-forming agent and provides a good model system for studying the interactions between metal ions and amphiphiles.  相似文献   

19.
The rhizotoxicity of aluminium at low-pH with Al(3+) and at high pH with Al(OH)-(4) as the main Al species was studied. Aluminium reduced root growth to similar levels at pH 8.0 and pH 4.3, although the mononuclear Al concentration at pH 8.0 was three times lower than at pH 4.3. Al contents of root apices were much higher at pH 8 than at pH 4.3. Callose was induced only marginally at pH 8 and the formation was confined to the epidermis, whereas it proceeded through the cortex with time at pH 4.3. Well-documented genotypical differences in callose formation and Al accumulation could not be found at pH 8. The largest fraction of the root-tip Al was recovered in the cell-wall fraction independent of the solution pH. A sequential extraction of isolated cell walls suggests that most of the cell-wall Al was precipitated Al(OH)(3) at pH 8.0. This can be explained by a drastic pH reduction in the root apoplastic sap to 6.2, whereas at bulk solution pH 4.3 it rose to 5.6. Al precipitation was also confirmed by the microscopic localization of Al. At pH 8, Al could mostly be found in the epidermis, but in the apoplast of the outer cortex at pH 4.3. It is proposed here that at pH 4.3, Al(3+) inhibits root growth through binding to sensitive binding sites in the apoplast of the epidermis and the outer cortex. At pH 8, Al(OH)(3) precipitation in the epidermis causes a mechanical barrier thus impairing the root-growth control of the epidermis.  相似文献   

20.
The effects of the transmembrane alpha-helical peptide Ac-K(2)(LA)(12)K(2)-amide ((LA)(12)) on the phase transition and dynamics of saturated dimyristoylphosphatidylcholine (DMPC) membranes were investigated at different pH using conventional and saturation-recovery EPR observations of phosphatidylcholine spin labels. At a peptide-to-DMPC ratio of 1/10, the main phase-transition temperature of the DMPC bilayer is decreased by 4.0 degrees C when measured at pH 7.0, by 1.6 degrees C when measured at pH 9.5, and not affected when measured at pH 11.5. This reversible pH effect is due to the subsequent neutralization of the positive charges of lysine side chains at both ends of (LA)(12). Apparent pK(a)s of the lysine side chain amino groups of (LA)(12) in DMPC bilayer are 8.6 and approximately 10.9, as compared with the pK(a) value of 10.5 for these groups when lysine is dissolved in water. Saturation-recovery curves as a function of oxygen concentration using phosphatidylcholine spin labels in DMPC bilayer containing (LA)(12) are always mono-exponential when measured at pH 7.0 and 9.5. This observation is consistent with the hypothesis that the lipid exchange rates among the bulk, boundary, and (LA)(12)-rich regions are faster than 0.5 micros, the electron spin-lattice relaxation time in the presence of molecular oxygen, suggesting that stable oligomers of (LA)(12) do not form. Neutralization of one lysine side chain positive charge on each end of the peptide significantly decreases the ordering effect of (LA)(12) on the lipid hydrocarbon chains, while its effect on the reorientational motion of terminal groups of lipid hydrocarbon chains is rather moderate. It does not affect the local diffusion-solubility product of oxygen measured in the DMPC-(LA)(12) membrane interior.  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