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.
SPARC (secreted protein, acidic and rich in cysteine, also known as osteonectin and BM-40) is a metal-binding glycoprotein secreted by a variety of cultured cells and characteristic of tissues undergoing morphogenesis, remodeling, and repair. Recently it has been shown that SPARC inhibits the progression of the endothelial cell cycle in mid-G1, and that a synthetic peptide (amino acids 54–73 of secreted murine SPARC, peptide 2.1) from a cationic, disulfide-bonded region was in part responsible for the growth-suppressing activity [Funk and Sage (1991): Proc Natl Acad Sci USA 88:2648–2652]. Moreover, SPARC was shown to interact directly with bovine aortic endothelial (BAE) cells through a C-terminal EF-hand sequence comprising a high-affinity Ca2+-binding site of SPARC and represented by a synthetic peptide (amino acids 254–273) termed 4.2 [Yost and Sage (1993): J Biol Chem 268:25790–25796]. In this study we show that peptide 4.2 is a more potent inhibitor of DNA synthesis that acts cooperatively with peptide 2.1 to diminish the incorporation of [3H]-thymidine by both BAE and bovine capillary endothelial (BCE) cells. At concentrations of 0.019–0.26 mM peptide 4.2, thymidine incorporation by BAE cells was decreased incrementally, relative to control values, from approximately 100 to 10%. Although somewhat less responsive, BCE cells exhibited a dose-responsive decrement in thymidine incorporation, with a maximal inhibition of 55% at 0.39 mM. The inhibitory effect of peptide 4.2 was essentially independent of heparin and basic fibroblast growth factor and was blocked by anti-SPARC peptide 4.2 IgG, but not by antibodies specific for other domains of SPARC. To identify residues that were necessary for inhibition of DNA synthesis, we introduced single amino acid substitutions into synthetic peptide 4.2 and tested their activities and cell-surface binding characteristics on endothelial cells. Two peptides displayed null to diminished effects in the bioassays that were concentration-dependent: peptide 4.2 K, containing an Asp258 → Lys substitution, and peptide 4.2 AA, in which the two disulfide-bonded Cys (positions 255 and 271) were changed to Ala residues. Peptide 4.2 K, which failed to fulfill the EF-hand consensus formula, exhibited an anomalous fluorescence emission spectrum, in comparison with the wild-type 4.2 sequence, that was indicative of a compromised affinity for Ca2+. Moreover, ablation of the disulfide bond in peptide 4.2 AA potentially destabilized the Ca2+-binding loop structure, as assessed by fluorescence spectroscopy, such that the peptide competed poorly for the binding of [125I]-peptide 4.2 to BAE cells. We conclude both that Ca2+-coordinating Asp at position 258 and the conformation of peptide 4.2 are necessary for the inhibition of DNA synthesis by SPARC in cultured endothelial cells.  相似文献   

2.
SPARC, a counteradhesive matricellular protein, inhibits endothelial cell adhesion and proliferation, but the pathways through which these activities are blocked are not known. In this study, we used inhibitors of major signaling proteins to identify mediators through which SPARC exerts its counteradhesive and antiproliferative functions. Pretreatments with the general protein tyrosine kinase (PTK) inhibitors, herbimycin A and genistein, protected against the inhibitory effect of SPARC on bovine aortic endothelial (BAE) cell spreading by more than 60 %. Similar pretreatments with PTK inhibitors significantly blocked the diminishment of focal adhesions by SPARC in confluent BAE cell monolayers, as determined by the formation of actin stress-fibers and the distribution of vinculin in focal adhesion plaques. Inhibition of endothelial cell cycle progression by SPARC and a calcium-binding SPARC peptide, however, was not affected by PTK inhibitors. Inhibition of DNA synthesis by SPARC was not reversed by inhibitors of the activity of protein kinase C (PKC), or of cAMP-dependent protein kinase (PKA), but was sensitive to pertussis (and to a lesser extent, cholera) toxin. The counteradhesive effect of SPARC on endothelial cells is, therefore, mediated through a tyrosine phosphorylation-dependent pathway, whereas its antiproliferative function is dependent, in part, on signal transduction via a G protein-coupled receptor. J. Cell. Biochem. 70:543–552, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

3.
《The Journal of cell biology》1990,111(6):3065-3076
Using synthetic peptides, we have identified two distinct regions of the glycoprotein SPARC (Secreted Protein Acidic and Rich in Cysteine) (osteonectin/BM-40) that inhibit cell spreading. One of these sites also contributes to the affinity of SPARC for extracellular matrix components. Peptides representing subregions of SPARC were synthesized and antipeptide antibodies were produced. Immunoglobulin fractions of sera recognizing an NH2-terminal peptide (designated 1.1) blocked SPARC- mediated anti-spreading activity. Furthermore, when peptides were added to newly plated endothelial cells or fibroblasts, peptide 1.1 and a peptide corresponding to the COOH terminal EF-hand domain (designated 4.2) inhibited cell spreading in a dose-dependent manner. These peptides exhibited anti-spreading activity at concentrations from 0.1 to 1 mM. The ability of peptides 1.1 and 4.2 to modulate cell shape was augmented by an inhibitor of protein synthesis and was blocked by specific antipeptide immunoglobulins. In addition to blocking cell spreading, peptide 4.2 competed for binding of [125I]SPARC and exhibited differential affinity for extracellular matrix molecules in solid-phase binding assays. The binding of peptide 4.2 to matrix components was Ca+(+)-dependent and displayed specificities similar to those of native SPARC. These studies demonstrate that both anti- spreading activity and affinity for collagens are functions of unique regions within the SPARC amino acid sequence. The finding that two separate regions of the SPARC protein contribute to its anti-spreading activity lead us to propose that multiple regions of the protein act in concert to regulate the interactions of cells with their extracellular matrix.  相似文献   

4.
SPARC, a Ca(2+)-binding glycoprotein that is expressed during tissue morphogenesis and functions as an inhibitor of cell spreading in vitro, was found to induce the secretion of an Mr = 45,000 protein in bovine aortic endothelial (BAE) cells. This protein was identified as type 1 plasminogen activator inhibitor (PAI-1) on Western blots with anti-PAI-1 antiserum. SPARC stimulated the secretion of PAI-1 protein into the medium of subconfluent BAE cells, but not confluent BAE cells, in a dose- and time-dependent manner. Secretion of PAI-1 into the culture medium was progressive and exhibited an increase of 3- to 7-fold over control values within 24 h after the addition of SPARC. Levels of PAI-1 mRNA were elevated 2-fold within 4 to 24 h after the addition of SPARC and did not increase with higher concentrations of SPARC. Since the induction of PAI-1 mRNA by SPARC was not blocked by cycloheximide, de novo protein synthesis was apparently not required for this stimulation. Control experiments showed that the induction of PAI-1 was not due to contamination of the SPARC preparations with endotoxin. These data demonstrate that SPARC induces the biosynthesis of PAI-1 in BAE cells and suggest a role for SPARC in the regulation of fibrinolysis and in the control of proteolytic events in remodeling tissues.  相似文献   

5.
Endothelioma cells expressing the polyoma virus middle T oncogene induced hemangiomas in mice by the recruitment of nonproliferating endothelial cells from host blood vessels (Williams et al. 1989). I now report that SPARC, a Ca(2+)-binding glycoprotein that perturbs cell-matrix interactions and inhibits the endothelial cell cycle, is produced by endothelioma cells and is in part responsible for the alterations in the morphology and growth that occur when nontransformed bovine aortic endothelial cells are cocultured with endothelioma cells. Normal endothelial cells cocultured with two different middle T-positive endothelial cell lines, termed End cells, exhibited changes in shape that were accompanied by the formation of cell clusters. Media conditioned by End cells repressed proliferation of normal endothelial cells, but enhanced that of an established line of murine capillary endothelium. Radiolabeling studies revealed no apparent differences in the profile of proteins secreted by aortic or capillary cells cultured in End cell conditioned media. Characterization of proteins produced by End cells led to the identification of type IV collagen, laminin, entactin, and SPARC as major secreted products. Although SPARC did not affect the morphology of End or capillary cells, it was associated with overt changes in the shape of aortic endothelial cells. Moreover, SPARC and a synthetic peptide from SPARC domain II inhibited the incorporation of [3H]thymidine by aortic cells, but had minimal to no effect on the capillary endothelial cell line. The inhibition of growth exhibited by aortic endothelial cells cultured in End cell conditioned media could be partially reversed by antibodies specific for SPARC and SPARC peptides. These studies indicate a potential role for SPARC in the generation of hemangiomas by End cells in vivo, a process that requires normal (host) endothelial cells to disengage from the extracellular matrix, withdraw from the cell cycle, migrate, and reassociate into the disorganized cellular networks that comprise cavernous and capillary hemangiomas.  相似文献   

6.
SPARC is a transiently expressed extracellular matrix-binding protein that alters cell shape and regulates endothelial cell proliferation in vitro. In this study, we show that SPARC mRNA and protein are synthesized by endothelial cells during angiogenesis in vivo. SPARC and peptides derived from a cationic region of the protein (amino acids 113- 130) stimulated the formation of endothelial cords in vitro; moreover, these peptides stimulated angiogenesis in vivo. Mapping of the active domain demonstrated that the sequence KGHK was responsible for most of the angiogenic activity; substitution of the His residue decreased the effect. We found that proteolysis of SPARC provided a source of KGHK, GHK, and longer peptides that contained these sequences. Although the Cu(2+)-GHK complex had been identified as a mitogen/morphogen in normal human plasma, we found KGHK and longer peptides to be potent stimulators of angiogenesis. SPARC113-130 and KGHK were shown to bind Cu2+ with high affinity; however, previous incubation with Cu2+ was not required for the stimulatory activity. Since a peptide from a second cationic region of SPARC (SPARC54-73) also bound Cu2+ but had no effect on angiogenesis, the angiogenic activity appeared to be sequence specific and independent of bound Cu2+. Thus, specific degradation of SPARC, a matrix-associated protein expressed by endothelial cells during vascular remodeling, releases a bioactive peptide or peptides, containing the sequence (K)GHK, that could regulate angiogenesis in vivo.  相似文献   

7.
Migration of endothelial cells is requisite to wound repair and angiogenesis. Since the glycoprotein SPARC (secreted protein, acidic and rich in cysteine) is associated with remodeling, cellular migration, and angiogenesis in vitro, we questioned whether SPARC might influence the motility of endothelial cells. In this study we show that, in the absence of serum, exogenous SPARC inhibits the migration of bovine aortic endothelial cells induced by bFGF. Similar results were obtained from two different assays, in which cell migration was measured in a Boyden chamber and in monolayer culture after an experimental wound. Without bFGF, the migration of endothelial cells was unaffected by SPARC. The inhibitory effect of SPARC on cell motility was dose-dependent, required the presence of Ca2+, was mimicked by synthetic peptides from the N- and C-terminal Ca(2+)-binding domains of the protein, and was not seen in the presence of serum. Modulation of the activities of secreted and cell-associated proteases, including plasminogen activators and metalloproteinases, appeared not to be responsible for the effects that we observed on the motility of endothelial cells. Moreover, a molecular interaction between SPARC and bFGF was not detected, and SPARC did not interfere with the binding of bFGF to high-affinity receptors on endothelial cells. Finally, in culture medium that contained serum, SPARC inhibited the incorporation of [3H]-thymidine into newly synthesized DNA, both in the absence and presence of bFGF. However, DNA synthesis was not affected by SPARC when the cells were plated on gelatin or fibronectin in serum-free medium. We propose that the combined action of a serum factor and SPARC regulates both endothelial cell proliferation and migration and coordinates these events during morphogenetic processes such as wound repair and angiogenesis.  相似文献   

8.
SPARC (Secreted Protein Acidic and Rich in Cysteine) is a Ca+2-binding glycoprotein that is differentially associated with morphogenesis, remodeling, cellular migration, and proliferation. We show here that exogenous SPARC, added to cells in culture, was associated with profound changes in cell shape, caused rapid, partial detachment of a confluent monolayer, and inhibited spreading of newly plated cells. Bovine aortic endothelial cells, exposed to 2-40 micrograms SPARC/ml per 2 x 10(6) cells, exhibited a rounded morphology in a dose-dependent manner but remained attached to plastic or collagen-coated surfaces. These round cells synthesized protein, uniformly excluded trypan blue, and grew in aggregates after replating in media without SPARC. SPARC caused rounding of bovine endothelial cells, fibroblasts, and smooth muscle cells; however, the cell lines F9, PYS-2, and 3T3 were not affected. The activity of native SPARC was inhibited by heat denaturation and prior incubation with anti-SPARC IgG. The effect of SPARC on endothelial cells appeared to be independent of the rounding phenomenon produced by the peptide GRGDSP. Immunofluorescence localization of SPARC on endothelial cells showed preferential distribution at the leading edges of membranous extensions. SPARC bound Ca+2 in both amino- and carboxyl-terminal (EF-hand) domains and required this cation for maintenance of native structure. Solid-phase binding assays indicated a preferential affinity of native SPARC for several proteins comprising the extracellular matrix, including types III and V collagen, and thrombospondin. This binding was saturable, Ca+2 dependent, and inhibited by anti-SPARC IgG. Endothelial cells also failed to spread on a substrate of native type III collagen complexed with SPARC. We propose that SPARC is an extracellular modulator of Ca+2 and cation-sensitive proteins or proteinases, which facilitates changes in cellular shape and disengagement of cells from the extracellular matrix.  相似文献   

9.
Cardiac tissue from mice that do not express secreted protein acidic and rich in cysteine (SPARC) have reduced amounts of insoluble collagen content at baseline and in response to pressure overload hypertrophy compared with wild-type (WT) mice. However, the cellular mechanism by which SPARC affects myocardial collagen is not clearly defined. Although expression of SPARC by cardiac myocytes has been detected in vitro, immunohistochemistry of hearts demonstrated SPARC staining primarily associated with interstitial fibroblastic cells. Primary cardiac fibroblasts isolated from SPARC-null and WT mice were assayed for collagen I synthesis by [(3)H]proline incorporation into procollagen and by immunoblot analysis of procollagen processing. Bacterial collagenase was used to discern intracellular from extracellular forms of collagen I. Increased amounts of collagen I were found associated with SPARC-null versus WT cells, and the proportion of total collagen I detected on SPARC-null fibroblasts without propeptides [collagen-α(1)(I)] was higher than in WT cells. In addition, the amount of total collagen sensitive to collagenase digestion (extracellular) was greater in SPARC-null cells than in WT cells, indicating an increase in cell surface-associated collagen in the absence of SPARC. Furthermore, higher levels of collagen type V, a fibrillar collagen implicated in collagen fibril initiation, were found in SPARC-null fibroblasts. The absence of SPARC did not result in significant differences in proliferation or in decreased production of procollagen I by cardiac fibroblasts. We conclude that SPARC regulates collagen in the heart by modulating procollagen processing and interactions with fibroblast cell surfaces. These results are consistent with decreased levels of interstitial collagen in the hearts of SPARC-null mice being due primarily to inefficient collagen deposition into the extracellular matrix rather than to differences in collagen production.  相似文献   

10.

Background

This laboratory previously analyzed the expression of SPARC in the parental UROtsa cells, their arsenite (As+3) and cadmium (Cd+2)-transformed cell lines, and tumor transplants generated from the transformed cells. It was demonstrated that SPARC expression was down-regulated to background levels in Cd+2-and As+3-transformed UROtsa cells and tumor transplants compared to parental cells. In the present study, the transformed cell lines were stably transfected with a SPARC expression vector to determine the effect of SPARC expression on the ability of the cells to form tumors in immune-compromised mice.

Methods

Real time PCR, western blotting, immunohistochemistry, and immunofluorescence were used to define the expression of SPARC in the As+3-and Cd+2-transformed cell lines, and urospheres isolated from these cell lines, following their stable transfection with an expression vector containing the SPARC open reading frame (ORF). Transplantation of the cultured cells into immune-compromised mice by subcutaneous injection was used to assess the effect of SPARC expression on tumors generated from the above cell lines and urospheres.

Results

It was shown that the As+3-and Cd+2-transformed UROtsa cells could undergo stable transfection with a SPARC expression vector and that the transfected cells expressed both SPARC mRNA and secreted protein. Tumors formed from these SPARC-transfected cells were shown to have no expression of SPARC. Urospheres isolated from cultures of the SPARC-transfected As+3-and Cd+2-transformed cell lines were shown to have only background expression of SPARC. Urospheres from both the non-transfected and SPARC-transfected cell lines were tumorigenic and thus fit the definition for a population of tumor initiating cells.

Conclusions

Tumor initiating cells isolated from SPARC-transfected As+3-and Cd+2-transformed cell lines have an inherent mechanism to suppress the expression of SPARC mRNA.  相似文献   

11.
Angiogenesis in vitro, the formation of capillary-like structures by cultured endothelial cells, is associated with changes in the expression of several extracellular matrix proteins. The expression of SPARC, a secreted collagen-binding glycoprotein, has been shown to increase significantly during this process. We now show that addition of purified SPARC protein, or an N-terminal synthetic peptide (SPARC4-23), to strains of bovine aortic endothelial cells undergoing angiogenesis in vitro resulted in a dose-dependent decrease in the synthesis of fibronectin and thrombospondin-1 and an increase in the synthesis of type 1-plasminogen activator inhibitor. SPARC decreased fibronectin mRNA by 75% over 48 h, an effect that was inhibited by anti-SPARC immunoglobulins. Levels of thrombospondin-1 mRNA were diminished by 80%. Over a similar time course, both mRNA and protein levels of type 1-plasminogen activator inhibitor (PAI-1) were enhanced by SPARC and the SPARC4-23 peptide. The effects were dose-dependent with concentrations of SPARC between 1 and 30 micrograms/ml. In contrast, no changes were observed in the levels of either type I collagen mRNA or secreted gelatinases. Half-maximal induction of PAI-1 mRNA or inhibition of fibronectin and thrombospondin mRNAs occurred with 2-5 micrograms/ml SPARC and approximately 0.05 mM SPARC4-23. Strains of endothelial cells that did not form cords and tubes in vitro had reduced or undetectable responses to SPARC under identical conditions. These results demonstrate that SPARC modulates the synthesis of a subset of secreted proteins and identify an N-terminal acidic sequence as a region of the protein that provides an active site. SPARC might therefore function, in part, to achieve an optimal ratio among different components of the extracellular matrix. This activity would be consistent with known effects of SPARC on cellular morphology and proliferation that might contribute to the regulation of angiogenesis in vivo.  相似文献   

12.
Objective: Adipocytes secrete various cytokines and matrix proteins. Several of them precipitate in obesity‐associated diseases, including atherosclerosis. In the current study, we have examined the expression of secreted protein, acidic and rich in cysteine (SPARC) in adipose tissue and its significance in obesity and coronary artery disease (CAD). Research Methods and Procedures: The SPARC mRNA expressions both in vivo and in vitro were detected by Northern blot analysis. Plasma SPARC concentrations were measured by enzyme immunosorbent assay. First, we investigated the plasma SPARC levels of 88 unrelated adult Japanese subjects (62 men and 26 women; average age: [± SD] 50 ± 12 years; body mass index [BMI]: 16 to 46 kg/m2). Additionally 31 subjects with CAD diagnosed by coronary angiography (20 men and 11 women) were also investigated. Results: Human adipose tissues expressed abundant SPARC mRNA. SPARC expression in adipose tissues was upregulated in obese db/db mice. Markedly enhanced expression of SPARC mRNA was observed in 3T3‐L1 fibroblasts during adipocyte differentiation. Consistent with these results, plasma SPARC levels proved a positive correlation with BMI in humans (r = 0.27; p < 0.01). Interestingly, plasma SPARC concentrations were significantly elevated in age‐ and BMI‐matched subjects with CAD (p < 0.05). Discussion: SPARC was expressed in adipose tissues and its expression was enhanced in obese mice. In human, plasma SPARC levels were elevated in obesity and CAD patients. This elevated SPARC may be involved in the progression of CAD.  相似文献   

13.
SPARC, a matricellular protein that affects cellular adhesion and proliferation, is produced in remodeling tissue and in pathologies involving fibrosis and angiogenesis. In this study we have asked whether peptides generated from cleavage of SPARC in the extracellular milieu can regulate angiogenesis. Matrix metalloproteinase (MMP)-3, but not MMP-1 or 9, showed significant activity toward SPARC. Limited digestion of recombinant human (rhu)SPARC with purified catalytic domain of rhuMMP-3 produced three major fragments, which were sequenced after purification by HPLC. Three synthetic peptides (Z-1, Z-2, and Z-3) representing motifs from each fragment were tested in distinct assays of angiogenesis. Peptide Z-1 (3.9 kDa, containing a Cu2+-binding sequence KHGK) exhibited a biphasic effect on [3H]thymidine incorporation by cultured endothelial cells and stimulated vascular growth in the chick chorioallantoic membrane (CAM). In contrast, peptides Z-2 (6.1 kDa, containing Ca2+-binding EF hand-1) and Z-3 (2.2 kDa, containing neither Cu2+-binding motifs nor EF hands), inhibited cell proliferation in a concentration-dependent manner and exhibited no effects on vessel growth in the CAM. Reciprocal results were obtained in a migration assay in native collagen gels: peptide Z-1 was ineffective over a range of concentrations, whereas Z-2 or Z-3 stimulated cell migration. Therefore, proteolysis of SPARC by MMP-3 produced peptides that regulate endothelial cell proliferation and/or migration in vitro in a mutually exclusive manner. One of these peptides containing KHGK also demonstrated a concentration-dependent effect on angiogenesis.  相似文献   

14.
SPARC (secreted protein, acidic and rich in cysteine) is an extracellular, Ca(2+)-binding protein that inhibits the spreading of newly plated cells and elicits a rounded morphology in spread cells. In this study, I investigated whether the rounding effect of SPARC depends on the ability of the protein to chelate Ca2+ at the cell surface. Bovine aortic endothelial cells were plated in the presence of different concentrations of SPARC and Ca2+; control experiments were performed with 1 mM EGTA and with Mg2+. Quantitative estimates of cell rounding were calculated according to a rounding index. SPARC, at concentrations between 0.15 and 0.58 microM, elicited rounding (or prevented spreading) of cells cultured for 16-38 h in 0.5-2.0 mM Ca2+. Addition of 0.5-2.0 mM Mg2+ to cells previously rounded in the presence of SPARC did not abrogate the effect of SPARC. When the levels of extracellular Ca2+ were adjusted with 1 mM EGTA to maximum values ranging from 7.1 to 320 microM, cells displayed a rounded morphology in the presence of exogenous SPARC. Although the rounding induced by 1 mM EGTA was essentially reversed by the inclusion of 2 mM Ca2+, cultures containing these reagents together with SPARC maintained the rounded phenotype. These results do not support a mechanism that involves the abstraction of Ca2+ from proteins at the cell surface or the provision of Ca2+ from native extracellular SPARC to cells. Therefore, SPARC does not appear to act as a local chelator of extracellular Ca2+ and Mg2+ and presumably exerts its function as a modulator of cell shape via a different pathway.  相似文献   

15.
The inhibition of human prostatic epithelial cell (MA-160) replication by cAMP and certain analogs was explored in tissue cultures. When untreated fetal bovine serum was used to supplement the culture medium, cyclic AMP (cAMP) markedly inhibited cell growth. The inhibition was reversed by equimolar concentrations of uridine. Inhibition by 8-methyl-thio-cAMP (MES) was somewhat less effective and was not reversed by uridine. After heat treatment of the fetal bovine serum, which inactivated the cAMP phosphodiesterases, cAMP became less effective in cell growth inhibition, whereas the activity of MES remained unaltered. Dibutyryl cAMP (db-cAMP) had no effect on cell growth, however, when combined with the phosphodiesterase inhibitor, 1-methyl-3-isobutylxanthine (MIX), significant retardation of cell replication was observed. Cells treated for 24 h with 0.5 mM MES took up and incorporated significantly less [3H]TdR and [3H]uridine than control cells. Treatment of cells with 0.5 mM cAMP for 24 h, on the other hand, resulted in both substantially increased [3H]TdR uptake and increased [3H]uridine incorporation into RNA. The effects of similar treatment with db-cAMP plus MIX closely paralleled those of MES with marked inhibition of the uptake and incorporation of both thymidine and uridine.  相似文献   

16.
Recent observations support an active role for the vascular endothelial cell in the induction and evolution of the inflammatory response. Since prior studies suggested that cultured bovine endothelial cells express high affinity binding sites for the neutrophil chemotactic oligopeptide formyl methionyl-leucyl-phenylalanine (f-Met-Leu-Phe), we sought to further characterize the interaction between formyl peptide chemoattractants and human vascular endothelial cells. Cultured human umbilical vein endothelial cells and peripheral blood neutrophils specifically bound f-Met-Leu-[3H]Phe, whereas specific binding to cultured fibroblasts, smooth muscle, and epithelial cells was negligible. Endothelial cells expressed 3.6 +/- 0.7 X 10(5) binding sites/cell with a Kd of 210 +/- 31 nM. Although the hexapeptide formyl norleucyl-leucyl-phenylalanyl-norleucyl-tyrosyl-lysine (f-Nle-Leu-Phe-Nle-Tyr-Lys) and the tetrapeptide f-Met-Leu-Phe-Lys completed with f-Met-Leu-[3H]Phe for binding to endothelial cells, specific binding of 125I-f-Nl-Leu-Phe-Tyr-Lys or f-Met-Leu-Phe-Lys-fluorescein to endothelial cells was not observed, suggesting that steric constraints on formyl peptide binding differ between endothelial cells and leukocytes. At 37 degrees C, cell-associated f-Met-Leu-[3H]Phe greatly exceeded that bound at 0 degrees C and was incorporated predominantly into a nondisplaceable compartment. Release of f-Met-Leu-[3H]Phe or radioactive breakdown products from this compartment was time- and temperature-dependent with a t1/2 of approximately equal to 20 min at 37 degrees C. Resolution of the radioactive products released from f-Met-Leu-[3H]Phe-loaded endothelial cells by thin layer chromatography indicated that greater than or equal to 57% of the released material co-migrated with intact f-Met-Leu-[3H]Phe. Degradative release was blocked by agents that interfere with lysosomal acidification. The radioactive material released from f-Met-Leu-[3H]Phe-loaded endothelial cells bound specifically to neutrophils. This binding was inhibited 50.2 +/- 6.4% by a greater than or equal to 10(3)-fold excess of nonradioactive f-Met-Leu-Phe whereas binding of authentic f-Met-Leu-[3H]Phe was inhibited 89.4 +/- 3.0%. Supernatant obtained from f-Met-Leu-[3H]Phe-loaded endothelial cells elicited a rise in neutrophil cytosolic free calcium ([Ca2+]i) measured by quin2 fluorescence. The change in neutrophil [Ca2+]i depended on ligand binding to the neutrophil formyl peptide receptor since endothelial supernatants were devoid of activity in the presence of the f-Met-Leu-Phe antagonist, tert-butoxycarbonyl-Phe-Leu-Phe-Leu-Phe.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
Abstract

The binding characteristics of [3H]quinuclidinyl benzilate ([3H]QNB) to isolated crude membranes of cultured bovine aortic endothelial cells were investigated. [3H]QNB bound to endothelial cell membranes with high affinity (kD = 0.056 nM) and limited capacity (132 fmol/mg DNA). The binding specificity, order of affinity and inhibition constants (Ki) were determined by displacement of bound [3H]QNB with unlabeled ligands. The order of affinity was QNB > atropine > 4-diphenylacetoxy-N-methyl-piperidine methiodide (4-DAMP) > p-fluoro-hexahydro-sila-difenidol (p-F-HHSiD) (M3 antagonist) > pirenzepine (M1 antagonist) > AFDX-116 (M2 antagonist) > (4-hydroxy-2-butynyl) trimethylammonium chloride m-chlorocarbanilate (McN-A-343, M1 agonist). These observations suggest that muscarinic receptors of endothelial cells in culture are likely to be of M3 and M1 subtype. Northern blot analysis of receptor subtypes using cDNA probes did not provide conclusive results due to the low level expression of these receptors in cultured cells. Solubilization of protein bound [3H]QNB with 1% digitonin and 0.02% cholate followed by analysis on sucrose density gradients demonstrated the presence of a specifically bound [3H]QNB-protein complex sedimenting at the 6.2S region of the gradient. These data demonstrate the presence of muscarinic acetylcholine receptor protein in cultured bovine aortic endothelial cells.  相似文献   

18.
Fetal rat lung fibroblasts characteristically increase their triacylglycerol (TG) stores during development. Both fibroblasts and alveolar type II (TII) cells can synthesize TG de novo, but only fibroblasts can absorb TG from culture medium, and retain the TG in a stable state. When fibroblasts pre-labelled with [3H]triolein are recombined with TII cells in organotypic culture the radiolabel appears in TII cell disaturated phosphatidylcholine (disatPC). When fibroblasts are preloaded with increasing amounts of TG there is a commensurate increase in TII cell disatPC following organotypic culture. Comparison of [3H]triacylglycerol and [14C]glucose incorporation into type II cell phospholipids revealed preferential use of TG for the surface-active phospholipids disatPC (10-fold greater) and phosphatidylglycerol (23-fold greater). These in vitro data suggest that fibroblasts provide lipid substrate for TII cell surfactant phospholipid synthesis.  相似文献   

19.
Selection and characterization of bovine aortic endothelial cells   总被引:17,自引:0,他引:17  
Summary This paper reports techniques for isolation, selection and long-term passage of bovine aortic endothelium (BAE). A [3H]thymidine-selection technique was developed to limit overgrowth of cultures by contaminating smooth-muscle cells. The resulting cultures could be passaged for a replicative life span of 35 to 40 doublings and maintained a stable, normal karyotype throughout this period. Despite the fact that these cultures reached a stable monolayer with density-inhibited growth state, postconfluent cells showed focal areas of a second growth pattern called “sprouting.” This was seen only when cultures were maintained at high densities for periods of 1 to 2 weeks. Ultrastructural analysis, as well as immunofluorescence studies with markers for endothelial cells (factor VIII) and smooth-muscle cells (actin), indicates that this phenomenon is not due to overgrowth of a residual population of smooth-muscle cells, but may represent a second growth pattern of the endothelial cells themselves. This research was supported by NIH Grant HL 18645. This work was done during the tenure of an Established Investigationship of the American Heart Association.  相似文献   

20.
A tridecapeptide containing tritium-labelled lysine and corresponding closely to residues 98 to 110 of the α chain of type I collagen was synthesized by the solid-phase method. Gly-Leu-Hyp-Gly-Nle-[4,5-3H]Lys-Gly-His-Arg-Gly-Phe-Ser-Gly was used as a substrate of human protocollagen lysyl hydroxylase (peptidyllysine, 2-oxoglutarate: oxygen 5-oxidoreductase, EC 1.14.11.4) obtained from dermal fibroblasts. L-[4,5-3H]Lysine was converted to Nα-t-butyloxycarbonyl-N?-o-chlorobenzyloxycarbonyl[3H]lysine which was incorporated during stepwise synthesis of the peptide. The chemical and radiochemical purities and specific activity of the completed peptide were characterized. A non-radiolabelled analogue of the peptide inhibited the hydroxylation of [3H]lysine-containing protocollagen by human lysyl hydroxylase, indicating that the synthetic peptide interacted with the enzyme. The peptide containing [3H]lysine was a substrate for lysyl hydroxylase and permitted direct measurement of enzyme activity in relatively crude cell extracts by a tritium-release assay. Extracts of cultured fibroblasts from a patient with an autosomal recessive pattern of inheritance for Ehlers-Danlos syndrome type VI had activities for tritium release from either the radiolabelled synthetic peptide or from [3H]lysine-containing protocollagen that were only 30% of those from control cells. These data indicate that a stable, well-defined synthetic peptide containing [3H]lysine is a useful substrate for studies of genetically variant lysyl hydroxylase from cultured human cells.  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