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.
Two small hydrophobic proteins, SP-B and SP-C, are responsible for rapid adsorption of pulmonary surfactant to the air/water interface. Despite their physiological importance, the number of protein molecules required to trigger an absorption event remains unknown. To investigate this issue, we varied the protein content of calf lung surfactant extract (CLSE) by dilution with protein-depleted surfactant lipids (neutral and phospholipids, N&PL). Vesicles of a constant size and of composition ranging between 100% N&PL and 100% CLSE were generated by probe sonication. Their adsorption kinetics to an air/water interface were monitored at different temperatures using a Wilhelmy plate to measure surface tension. When plotted versus protein concentration, the adsorption rates during the initial change in surface tension exhibit a diphasic behavior, first increasing rapidly and linearly between 0% and 25% CLSE, and then more slowly at higher concentrations. Direct linearity at low protein content (0-5% CLSE ratio) was confirmed at 37 degrees C. These observations argue against cooperative behavior, for which the adsorption rate would first rise slowly with the protein content, and then increase suddenly once the critical number of proteins on each vesicle is reached. The apparent activation energy E(a) and the free energy of activation DeltaG(0)*, calculated from the temperature dependence of adsorption, further support the view that at least the early stages of protein-induced surfactant adsorption proceeds through a sequence of events involving not several, but a single surfactant protein.  相似文献   

2.
To determine how different constituents of pulmonary surfactant affect its phase behavior, we measured wide-angle x-ray scattering (WAXS) from oriented bilayers. Samples contained the nonpolar and phospholipids (N&PL) obtained from calf lung surfactant extract (CLSE), which also contains the hydrophobic surfactant proteins SP-B and SP-C. Mixtures with different ratios of N&PL and CLSE provided the same set of lipids with different amounts of the proteins. At 37°C, N&PL by itself forms coexisting Lα and Lβ phases. In the Lβ structure, the acyl chains of the phospholipids occupy an ordered array that has melted by 40°C. This behavior suggests that the Lβ composition is dominated by dipalmitoyl phosphatidylcholine (DPPC), which is the most prevalent component of CLSE. The Lβ chains, however, lack the tilt of the Lβ phase formed by pure DPPC. At 40°C, WAXS also detects an additional diffracted intensity, the location of which suggests a correlation among the phospholipid headgroups. The mixed samples of N&PL with CLSE show that increasing amounts of the proteins disrupt both the Lβ phase and the headgroup correlation. With physiological levels of the proteins in CLSE, both types of order are absent. These results with bilayers at physiological temperatures indicate that the hydrophobic surfactant proteins disrupt the ordered structures that have long been considered essential for the ability of pulmonary surfactant to sustain low surface tensions. They agree with prior fluorescence micrographic results from monomolecular films of CLSE, suggesting that at physiological temperatures, any ordered phase is likely to be absent or occupy a minimal interfacial area.  相似文献   

3.
Schram V  Hall SB 《Biophysical journal》2001,81(3):1536-1546
We determined the influence of the two hydrophobic proteins, SP-B and SP-C, on the thermodynamic barriers that limit adsorption of pulmonary surfactant to the air-water interface. We compared the temperature and concentration dependence of adsorption, measured by monitoring surface tension, between calf lung surfactant extract (CLSE) and the complete set of neutral and phospholipids (N&PL) without the proteins. Three stages generally characterized the various adsorption isotherms: an initial delay during which surface tension remained constant, a fall in surface tension at decreasing rates, and, for experiments that reached approximately 40 mN/m, a late acceleration of the fall in surface tension to approximately 25 mN/m. For the initial change in surface tension, the surfactant proteins accelerated adsorption for CLSE relative to N&PL by more than ten-fold, reducing the Gibbs free energy of transition (DeltaG(O)) from 119 to 112 kJ/mole. For the lipids alone in N&PL, the enthalpy of transition (DeltaH(O), 54 kJ/mole) and entropy (-T. DeltaS, 65 kJ/mole at 37 degrees C) made roughly equal contributions to DeltaG(O). The proteins in CLSE had little effect on -T. DeltaS(O) (68 kJ/mole), but lowered DeltaG(O) for CLSE by reducing DeltaH(O) (44 kJ/mole). Models of the detailed mechanisms by which the proteins facilitate adsorption must meet these thermodynamic constraints.  相似文献   

4.
These in vitro experiments study a potential mechanism by which plasma proteins, found in the alveoli during pulmonary edema and hemorrhage, may act to inhibit the surface activity of pulmonary surfactant. The results indicate that the inhibition of the adsorption facility and surface tension lowering ability of a calf lung surfactant extract (CLSE) by albumin, hemoglobin, or fibrinogen may be completely abolished by centrifugation of the protein-surfactant mixture at 12,500 x g. Furthermore, albumin, hemoglobin and fibrinogen (1.25 mg/ml) were shown to inhibit the adsorption of high concentrations of CLSE (0.32 mg/ml), normally unaffected by the addition of exogenous proteins, when the CLSE was injected into the subphase under a preformed protein surface film. Similarly, injection of large amounts of these proteins (2.5 mg/ml) into the subphase beneath a preformed CLSE surface film was without effect, even though the CLSE concentration was only 0.06 mg/ml, a surfactant concentration which is normally inhibited by even small amounts of exogenous protein. Taken together, the data suggest that some proteins may inhibit surfactant function by preventing the surfactant phospholipids from adsorbing to the air-liquid interface, possibly by a competition between the proteins and CLSE phospholipids for space at the air-liquid interface rather than direct molecular interactions between proteins and surfactant.  相似文献   

5.
Surface activity and sensitivity to inhibition from phospholipase A2 (PLA2), lysophosphatidylcholine (LPC), and serum albumin were studied for a synthetic C16:0 diether phosphonolipid (DEPN-8) combined with 1.5% by weight of mixed hydrophobic surfactant proteins (SP)-B/C purified from calf lung surfactant extract (CLSE). Pure DEPN-8 had better adsorption and film respreading than the major lung surfactant phospholipid dipalmitoyl phosphatidylcholine and reached minimum surface tensions <1 mN/m under dynamic compression on the Wilhelmy balance and on a pulsating bubble surfactometer (37 degrees C, 20 cycles/min, 50% area compression). DEPN-8 + 1.5% SP-B/C exhibited even greater adsorption and had overall dynamic surface tension lowering equal to CLSE on the bubble. In addition, films of DEPN-8 + 1.5% SP-B/C on the Wilhelmy balance had better respreading than CLSE after seven (but not two) cycles of compression-expansion at 23 degrees C. DEPN-8 is structurally resistant to degradation by PLA2, and DEPN-8 + 1.5% SP-B/C maintained high adsorption and dynamic surface activity in the presence of this enzyme. Incubation of CLSE with PLA2 led to chemical degradation, generation of LPC, and reduced surface activity. DEPN-8 + 1.5% SP-B/C was also more resistant than CLSE to direct biophysical inhibition by LPC, and the two were similar in their sensitivity to biophysical inhibition by serum albumin. These findings indicate that synthetic surfactants containing DEPN-8 combined with surfactant proteins or related synthetic peptides have potential utility for treating surfactant dysfunction in inflammatory lung injury.  相似文献   

6.
This research studies the biophysical surface activity of synthetic phospholipids combined in vitro with purified lung surfactant apoprotein, having an Mr of 6000. Hydrophobic surfactant-associated protein (SAP-6) was delipidated and purified from both bovine and canine lung lavage, and was combined in vitro with a synthetic phospholipid mixture (SM) of similar composition to natural lung surfactant phospholipids. SM phospholipids were also combined and studied biophysically with another purified surfactant-associated protein, SAP-35. The biophysical activity of synthetic phospholipid-apoprotein combinants was assessed by measurements of adsorption facility and dynamic surface tension lowering ability at 37 degrees C. The SM-SAP-6 combinants had adsorption facility equivalent to natural lung surfactant, and to the surfactant extract preparations CLSE and surfactant-TA used in exogenous surfactant replacement therapy for the neonatal Respiratory Distress Syndrome (RDS). The synthetic phospholipid-SAP-6 combinants also lowered surface tension to less than 1 dyne/cm under dynamic compression in an oscillating bubble apparatus at concentrations as low as 0.5 mg phospholipid/ml. A striking finding was that this excellent dynamic surface activity was preserved as SAP-6 composition was reduced to values as low as 5 micrograms/5 mg SM phospholipid (0.1% SAP-6 protein), an order of magnitude less than the 1% protein content of CLSE and surfactant-TA. Mixtures of SM phospholipids plus SAP-35, the major surfactant glycoprotein, had significantly lower biophysical activity, which did not approach that of a functional lung surfactant. These results suggest that synthetic exogenous surfactants of potential utility for replacement therapy in RDS can be formulated by combining synthetic phospholipids in vitro with specifically purified, hydrophobic surfactant-associated protein, SAP-6.  相似文献   

7.
The content-dependent activity of surfactant protein (SP)-B was studied in mixtures with dipalmitoyl phosphatidylcholine (DPPC), synthetic lipids (SL), and purified phospholipids (PPL) from calf lung surfactant extract (CLSE). At fixed SP-B content, adsorption and dynamic surface tension lowering were ordered as PPL/SP-B approximately SL/SP-B > DPPC/SP-B. All mixtures were similar in having increased surface activity as SP-B content was incrementally raised from 0.05 to 0.75% by weight. SP-B had small but measurable effects on interfacial properties even at very low levels < or =0.1% by weight. PPL/SP-B (0.75%) had the highest adsorption and dynamic surface activity, approaching the behavior of CLSE. All mixtures containing 0.75% SP-B reached minimum surface tensions <1 mN/m in pulsating bubble studies at low phospholipid concentration (1 mg/ml). Mixtures of PPL or SL with SP-B (0.5%) also had minimum surface tensions <1 mN/m at 1 mg/ml, whereas DPPC/SP-B (0.5%) reached <1 mN/m at 2.5 mg/ml. Physiological activity also was strongly dependent on SP-B content. The ability of instilled SL/SP-B mixtures to improve surfactant-deficient pressure-volume mechanics in excised lavaged rat lungs increased as SP-B content was raised from 0.1 to 0.75% by weight. This study emphasizes the crucial functional activity of SP-B in lung surfactants. Significant differences in SP-B content between exogenous surfactants used to treat respiratory disease could be associated with substantial activity variations.  相似文献   

8.
SP-B and SP-C alter diffusion in bilayers of pulmonary surfactant   总被引:2,自引:0,他引:2       下载免费PDF全文
Schram V  Hall SB 《Biophysical journal》2004,86(6):3734-3743
The hydrophobic proteins SP-B and SP-C promote rapid adsorption of pulmonary surfactant to an air/water interface by an unknown mechanism. We tested the hypothesis that these proteins accelerate adsorption by disrupting the structure of the lipid bilayer, either by a generalized increase in fluidity or by a focal induction of interfacial boundaries within the bilayer. We used fluorescence recovery after photobleaching to measure diffusion of nitrobenzoxadiazolyl-dimyristoyl-phosphatidylethanolamine between 11 and 54 degrees C in multilayers containing the complete set of lipids and proteins in calf lung surfactant extract (CLSE), or the complete set of neutral and phospholipids without the proteins. Above 35 degrees C, Arrhenius plots of diffusion were parallel for CLSE and neutral and phospholipids, but shifted to lower values for CLSE, suggesting that the proteins rigidify the lipid bilayer rather than producing the proposed increase in membrane fluidity. The slopes of the Arrhenius plots for CLSE were steeper below 35 degrees C, suggesting that the proteins induce phase separation at that temperature. The mobile fraction fell below 27 degrees C, consistent with a percolation threshold of coexisting gel and liquid-crystal phases. The induction of lateral phase separation in CLSE, however, does not correlate with apparent changes in adsorption kinetics at this temperature. Our results suggest that SP-B and SP-C accelerate adsorption through a mechanism other than the disruption of surfactant bilayers, possibly by stabilizing a high-energy, highly curved adsorption intermediate.  相似文献   

9.
Rapid adsorption of surfactant material to the air/liquid interface of the lung is essential for maintaining normal lung function. The detailed mechanism of this process, however, remains unclear. In this study, we elucidate the influence of lipid saturation grade and headgroup charge of surface layer lipids on surfactant protein (SP)-induced vesicle insertion into monolayers spread at the air/water interface of a film balance. We used dipalmitoylphosphatidlycholine (DPPC),1,2-dipalmitoyl-sn-glycero-3-phosphoglycerol (DPPG), 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC), and 1-palmitoyl-2-oleoyl-sn-glycero-3-phosphoglycerol (POPG) as monolayer lipids doped with either hydrophobic surfactant-specific protein SP-B or SP-C (0.2 and 0.4 mol %, respectively). Vesicles consisting of DPPC/DPPG (4:1, mol ratio) were injected into a stirred subphase to quantify adsorption kinetics. Based on kinetic film balance and fluorescence measurements, a refined model describing distinct steps of vesicle adsorption to surfactant monolayers is presented. First, in a protein-independent step, lipids from vesicles bridged to the interfacial film by Ca2+ ions are inserted into defects of a disordered monolayer at low surface pressures. Second, in a SP-facilitated step, active material insertion involving an SP-B- or SP-C-induced flip-flop of lipids occurs at higher surface pressures. Negatively charged lipids obviously influence the threshold pressures at which this second protein-mediated adsorption mechanism takes place.  相似文献   

10.
This study examines the direct inhibitory effects of Pneumocystis carinii (Pc) organisms and chemical components on the surface activity and composition of whole calf lung surfactant (WLS) and calf lung surfactant extract (CLSE) in vitro. Incubation of WLS suspensions with intact Pc organisms (10(7) per milligram of surfactant phospholipid) did not significantly alter total phospholipid levels or surfactant protein A content. Incubation with intact Pc organisms also did not impair dynamic surface tension lowering in suspensions of WLS or centrifuged large surfactant aggregates on a bubble surfactometer (37 degrees C, 20 cycles/min, 0.5 and 2.5 mg phospholipid/ml). However, exposure of WLS or CLSE to disrupted (sonicated) Pc organisms led to severe detriments in activity, with minimum surface tensions of 17-19 mN/m vs. <1 mN/m for surfactants alone. Extracted hydrophobic chemical components from Pc (98.8% lipids, 0.1 mM) reduced the surface activity of WLS and CLSE similarly to sonicated Pc organisms, whereas extracted hydrophilic chemical components from Pc (primarily proteins) had only minor effects on surface tension lowering. These results indicate that in addition to surfactant dysfunction induced by inflammatory lung injury and edema-derived inhibitors in Pc pneumonia, disrupted Pc organisms in the alveolar lumen also have the potential to directly inhibit endogenous and exogenous lung surfactants in affected patients.  相似文献   

11.
These experiments characterize the effects of hemoglobin and erythrocyte membrane lipids on the dynamic surface activity and adsorption facility of whole lung surfactant (LS) and a calf lung surfactant extract (CLSE) used clinically in surfactant replacement therapy for the neonatal respiratory distress syndrome (RDS). The results show that, at concentrations from 25 to 200 mg/ml, hemoglobin (Hb) increased the minimum dynamic surface tension of LS or CLSE mixtures (0.5 and 1.0 mumol/ml) from less than 1 to 25 dyn/cm on an oscillating bubble apparatus at 37 degrees C. Similarly, erythrocyte membrane lipids (0.5-3 mumol/ml) also prevented LS and CLSE suspensions (0.5-2.0 mumol/ml) from lowering surface tension below 19 dyn/cm under dynamic compression on the bubble. Surface pressure-time adsorption isotherms for LS suspensions (0.084 and 0.168 mumol phospholipid/ml) were also adversely affected by Hb (0.3-2.5 mg/ml), having a slower adsorption rate and magnitude. Significantly, these inhibitory effects of Hb and membrane lipids could be abolished if LS and CLSE concentrations were raised to high levels. In complementary physiological experiments, instillation of Hb, membrane lipids, or albumin into excised rat lungs was shown to cause a decrease in pressure-volume compliance. This decreased compliance was most prominent in lungs made partially surfactant deficient before inhibitor delivery and could be reversed by supplementation with active exogenous surfactant. Taken together, these data show that molecular components in hemorrhagic pulmonary edema can biophysically inactivate endogenous LS and adversely affect lung mechanics. Moreover, exogenous surfactant replacement can reverse this process even in the continued presence of inhibitor molecules and thus has potential utility in therapy for adult as well as neonatal RDS.  相似文献   

12.
The effects of surfactant protein (SP)-A on the dynamic surface tension lowering and resistance to inhibition of dispersions of calf lung surfactant extract (CLSE) and mixtures of synthetic phospholipids combined with SP-B,C hydrophobic apoproteins were studied at 37 degrees C and rapid cycling rate (20 cycles/min). Addition of SP-A to CLSE, which already contains SP-B and -C, gave a slight improvement in the time course of surface tension lowering on an oscillating bubble apparatus in the absence of inhibitory protein molecules such as albumin or hemoglobin. However, when these proteins were present at concentrations of 10-50 mg/ml, SP-A substantially improved the resistance of CLSE to their inhibitory effects. The beneficial effect of SP-A required the presence of Ca2+ ions, and disappeared when EDTA was substituted for this divalent cation in the subphase. The effect was also retained when SP-A was heated to 50 degrees C prior to addition to CLSE, but was abolished by heating SP-A to 99 degrees C. Additional studies showed that similar improvements in resistance to inhibition were found when SP-A was added to synthetic mixtures of dipalmitoyl phosphatidylcholine (DPPC):egg phosphatidylglycerol (PG) (80:20 by weight) reconstituted with 1% SP-B or SP-B and -C, but not to phospholipid mixtures containing only SP-C. The requirements for SP-B and calcium for the beneficial effects of SP-A on surface activity suggest that the formation of ordered, larger phospholipid-apoprotein aggregates may be involved in the process. The finding that SP-A enhances the ability of CLSE and other surfactant mixtures containing SP-B to resist inhibition is an advantage that will need to be weighed against other factors such as increased antigenicity and heat sensitivity in therapeutic applications in surfactant replacement therapy.  相似文献   

13.
The primary function of pulmonary surfactant is to reduce the surface tension (ST) created at the air–liquid interface in the lung. Surfactant is a complex mixture of lipids and proteins and its function is influenced by physiological parameters such as metabolic rate, body temperature and breathing. In the microchiropteran bat Chalinolobus gouldii these parameters fluctuate throughout a 24 h period. Here we examine the surface activity of surfactant from warm–active and torpid bats at both 24°C and 37°C to establish whether alterations in surfactant composition correlate with changes in surface activity. Bats were housed in a specially constructed bat room at Adelaide University, at 24°C and on a 8:16 h light:dark cycle. Surfactant was collected from bats sampled during torpor (25<Tb<28°C), and while active (Tb>35°C). Alterations in the lipid composition of surfactant occur with changes in the activity cycle. Most notable is an increase in surfactant cholesterol (Chol) with decreases in body temperature [Codd et al., Physiol. Biochem. Zool. 73 (2000) 605–612]. Surfactant from active bats was more surface active at higher temperatures, indicated by lower STmin and less film area compression required to reach STmin at 37°C than at 24°C. Conversely, surfactant from torpid bats was more active at lower temperatures, indicated by lower STmin and less area compression required to reach STmin at 24°C than at 37°C. Alterations in the Chol content of bat surfactant appear to be crucial to allow it to achieve low STs during torpor.  相似文献   

14.
This study focused on two hydrophobic fractions (HF-A and HF-B) isolated from porcine lung surfactant (LS) that had similar phospholipid composition, but HF-A consisted of the hydrophobic LS specific proteins (SP-B and SP-C), in contrast to HF-B. Monolayers spread in a Langmuir trough were formed at the air/water interface of both fractions and the rate of adsorption-desorption and the respreading potential of the LS constituents was studied during six consecutive compression/decompression cycles of the monolayers. By drawing a comparison between the behavior of HF-A and HF-B monolayers on the subphase of 150 mm NaCl, either with or without additional Ca2+, we estimated the role of hydrophobic LS proteins and Ca2+ ions for LS surface activity. The results demonstrated much higher ability of the HF-A sample, compared to HF-B, to maintain lower surface tension (γ) during monolayer compression and its better respreading capacity during decompression. For instance, at a surface concentration corresponding to 80 Å2 per phospholipid molecule, the HF-A monolayers showed a much lower γ max value (surface tension at 100% of the trough area), being ca. 31.0 mN/m, compared to the HF-B monolayers (γ max? 62.0 mN/m). The surface tension after compression to 20% of the initial area (γ min) reached ca. 7.0 and 19.0 mN/m in the HF-A and HF-B monolayers, respectively. Better respreading of the HF-A monolayers compared to the HF-B monolayers was due to the faster adsorption and spreading of LS phospholipids during decompression, facilitated by the hydrophobic proteins. As the phospholipid composition of both fractions was similar, we showed that the hydrophobic surfactant proteins were responsible also for the prevention of the irreversible loss of material from the surface during monolayer compression/decompression. The effects observed demonstrated also that the hydrophobic surfactant proteins were the stronger determinant, compared with Ca2+ ions, for the surface tension decrease and respreading of the monolayers during film compression/decompression. For instance, when the HF-A monolayers were spread on a subphase with an additional 5 mm Ca2+ ion content, no significant changes were detected in the γ min and γ max values between the first and sixth cycle, compared to the monolayers spread on a subphase of 150 mm NaCl only. However, in the absence of positively charged SP-B and SP-C (HF-B sample) in highly compressed monolayers, Ca2+ ions were able to cause the effects shown by SP-B and SP-C, although to a less extent. The role of the electrostatic and hydrophobic interactions is discussed for the better respreading of LS components in the presence of LS proteins and Ca2+ ions.  相似文献   

15.
Prior studies suggest that the hydrophobic surfactant proteins, SP-B and SP-C, promote adsorption of the lipids in pulmonary surfactant to an air-water interface by stabilizing a negatively curved rate-limiting structure that is intermediate between bilayer vesicles and the surface film. This model predicts that other peptides capable of stabilizing negative curvature should also promote lipid adsorption. Previous reports have shown that under appropriate conditions, gramicidin-A (GrA) induces dioleoyl phosphatidylcholine (DOPC), but not dimyristoyl phosphatidylcholine (DMPC), to form the negatively curved hexagonal-II (HII) phase. The studies reported here determined if GrA would produce the same effects on adsorption of DMPC and DOPC that the hydrophobic surfactant proteins have on the surfactant lipids. Small angle X-ray scattering and 31P-nuclear magnetic resonance confirmed that at the particular conditions used to study adsorption, GrA induced DOPC to form the HII phase, but DMPC remained lamellar. Measurements of surface tension showed that GrA in vesicles produced a general increase in the rate of adsorption for both phospholipids. When restricted to the interface, however, in preexisting films, GrA with DOPC, but not with DMPC, replicated the ability of the surfactant proteins to promote adsorption of vesicles containing only the lipids. The correlation between the structural and functional effects of GrA with the two phospholipids, and the similar effects on adsorption of GrA with DOPC and the hydrophobic surfactant proteins with the surfactant lipids fit with the model in which SP-B and SP-C facilitate adsorption by stabilizing a rate-limiting intermediate with negative curvature.  相似文献   

16.
Reported associations between air pollution and pulmonary and cardiovascular diseases prompted studies on the effects of gold nanoparticles (Au NP) on pulmonary surfactant function. Low levels (3.7 mol % Au/lipid, 0.98% wt/wt) markedly inhibited adsorption of a semisynthetic pulmonary surfactant (dipalmitoyl-phosphatidylcholine (DPPC)/palmitoyl-oleoyl-phosphatidylglycerol/surfactant protein B (SP-B); 70:30:1 wt %). Au NP also impeded the surfactant's ability to reduce surface tension (γ) to low levels during film compression and to respread during film expansion. Transmission electron microscopy showed that Au NP generated by a seed-growth method were spherical with diameters of ∼15 nm. Including palmitoyl-oleoyl-phosphatidylglycerol appeared to coat the NP with at least one lipid bilayer but did not affect NP shape or size. Similar overall observations occurred with dimyristoyl phosphatidylglycerol. Dipalmitoyl-phosphatidylglycerol was less effective in NP capping, although similar sized NP were formed. Including SP-B (1% wt/wt) appears to induce the formation of elongated strands of interacting threads with the fluid phosphatidylglycerols (PG). Including DPPC resulted in formation of aggregated, less spherical NP with a larger size distribution. With DPPC, strand formation due to SP-B was not observed. Agarose gel electrophoresis studies demonstrated that the aggregation induced by SP-B blocked migration of PG-coated NP. Migration was also influenced by the fluidity of the PGs. It is concluded that Au NP can interact with and sequester pulmonary surfactant phospholipids and, if inhaled from the atmosphere, could impede pulmonary surfactant function in the lung.  相似文献   

17.
In the current work we study with monolayer tensiometry and Brewster angle microscopy (BAM) the surface properties of Dipalmitoleoylphosphatidylethanolamine (DPoPE) films at the air/water interface in presence and absence of specific surfactant protein C (SP-C). DPoPE is used, as it readily forms both lamellar (Lα) and non-lamellar inverted hexagonal (HII) phases and appears as a suitable model phospholipid for probing the interfacial properties of distinct lipid phases. At pure air/water interface Lα shows faster adsorption and better surface disintegration than HII phase. The interaction of DPoPE molecules with SP-C (predeposited at the interface) results in equalizing of the interfacial disintegration of the both phases (reaching approximately the same equilibrium surface tension) although the adsorption kinetics of the lamellar phase remains much faster. Monolayer compression/decompression cycling revealed that the effect of SP-C on dynamic surface tensions (γ max and γ min) of mixed films is remarkably different for the two phases. If γ max for Lα decreased from the first to the third cycle, the opposite effect is registered for HII where γ max increases during cycling. Also the significant decrease of γ min for Lα in SP-C presence is not observed for HII phase. BAM studies reveal the formation of more uniform and homogeneously packed DPoPE monolayers in the presence of SP-C.  相似文献   

18.
To remove endotoxins (lipopolysaccharides; LPS) from cell products used as drugs, water-insoluble poly(ε-lysine) (PL) particles were prepared by cross-linking with PL originating from Streptomyces albulus and chloromethyloxirane (CMO). The apparent pKa (pKa,app) and the anion-exchange capacity of the particles were easily adjusted by changing the PL ratio and the CMO ratio. The higher the pKa,app, the greater the LPS-adsorption capacity of the particles. On the other hand, when the PL ratio (in the particles) increased to 75 unit-mol% or higher, the adsorption of bovine serum albumin by the particles also increased, but decreased with increasing ionic strength of the buffer to μ=0.2 or higher. The adsorption of γ-globulin increased with decreasing PL ratio to 65 unit-mol% or lower. As a result, when the PL ratio was 70 unit-mol% and the pKa,app was 6.7, the PL/CMO particles selectively removed LPS from various protein solutions that were naturally contaminated with LPS, at pH 6.0 and μ=0.05.  相似文献   

19.
Pharmacy-grade exogenous lung surfactant preparations of bovine and porcine origin, dispersed in physiological electrolyte solution have been studied. The organization and dynamics at the air/water interface at physiological temperature was analysed by neutron reflection. The results show that a well-defined surface phase is formed, consisting of a multilayer structure of lipid/protein bilayers alternating with aqueous layers, with a repetition period of about 70 Å and correlation depths of 3 to > 25 bilayers, depending on electrolyte composition and time. The experimental surfactant concentration of 0.15% (w/w) is far below that used in therapeutic application of exogenous surfactants and it is therefore likely that similar multilayer structures are also formed at the alveolar surface in the clinical situation during surfactant substitution therapy. Lung surfactant preparations in dry form swell in aqueous solution towards a limit of about 60% (w/w) of water, forming a lamellar liquid-crystalline phase above about 34 °C, which disperses into lamellar bodies at higher water concentrations. The lamellar spacings in the surface multilayers at the air/water interface are smaller than those in the saturated limit even though they are in contact with much greater water concentrations. The surface multilayers are laterally disordered in a way that is consistent with fragments of Lα-phase lamellae. The near surface layers of the multilayer structure have a significant protein content (only SP-B and SP-C are present in the preparations). The results demonstrate that a multilayer structure can be formed in exogenous surfactant even at very low concentrations and indicate that multilayers need to be incorporated into present interpretations of in vitro studies of similar lung surfactant preparations, which are largely based on monolayer models.  相似文献   

20.
Two DCM mutants (E40K and E54K) of tropomyosin (Tm) were examined using the thin-filament extraction/reconstitu­tion technique. The effects of the Ca2+, ATP, phos­phate (Pi), and ADP concentrations on isometric tension and its transients were studied at 25°C, and the results were com­pared to those for the WT protein. Our results indicate that both E40K and E54K have a significantly lower T HC (high Ca2+ ten­sion at pCa 4.66) (E40K: 1.21±0.06 T a, ±SEM, N = 34; E54K: 1.24±0.07 T a, N = 28), a significantly lower T LC (low- Ca2+ tension at pCa 7.0) (E40K: 0.07±0.02 T a, N = 34; E54K: 0.06±0.02 T a, N = 28), and a significantly lower T act (Ca2+ activatable tension) (T act = T HC–TLC, E40K: 1.15±0.08 T a, N = 34; E54K: 1.18±0.06 T a, N = 28) than WT (T HC = 1.53±0.07 T a, T LC = 0.12±0.01 T a, T act = 1.40±0.07 T a, N = 25). All tensions were normalized to T a ( = 13.9±0.8 kPa, N = 57), the ten­sion of actin-filament reconstituted cardiac fibers (myocardium) under the standard activating conditions. The Ca2+ sensitivity (pCa50) of E40K (5.23±0.02, N = 34) and E54K (5.24±0.03, N = 28) was similar to that of the WT protein (5.26±0.03, N = 25). The cooper­a­tivity increased significantly in E54K (3.73±0.25, N = 28) compared to WT (2.80±0.17, N = 25). Seven kinetic constants were deduced using sinusoidal analysis at pCa 4.66. These results enabled us to calculate the cross-bridge distribution in the strongly attached states, and thereby deduce the force/cross-bridge. The results indicate that the force/cross-bridge is ∼15% less in E54K than WT, but remains similar to that of the WT protein in the case of E40K. We conclude that over-inhibition of the actomyosin interaction by E40K and E54K Tm mutants leads to a decreased force-generating ability at systole, which is the main mechanism underlying the early pathogenesis of DCM.  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