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.
The present study for the first time evaluated both the in vitro and in vivo phosphorylation regions of bone sialoprotein (BSP) by utilizing multiple approaches and techniques. The in vitro phosphorylation sites were determined by 32P-labeling of native BSP using purified casein kinase II (CKII), followed by peptide mapping and solid-phase N-terminal sequence analyses. The in vivo phosphorylation sites were determined by (i) derivatization with 1-S-[14C]carboxymethyl-dithiothreitol ([14C] CM-DTT) of the proteolytic digests of BSP, followed by isolation and N-terminal peptide sequence analysis; and (ii) analyzing the proteolytic peptides of native BSP using matrix-assisted laser desorption/ionization-time of flight-mass spectrometry (MALDI-TOF-MS). Native BSP incorporated approximately 2.5 mol of phosphate/mol of BSP by CKII, which were distributed over four major peptide peaks and three shoulder peaks within the peptide map with varying degrees of phosphorylation. Further studies using the [14C] CM-DTT thiol reagent indicated that native and deglycosylated BSP incorporated 5.84 and 5.80 mol of 14C/mol of BSP, respectively. This confirmed that there were approximately 5.8 mol P-Ser/mol of BSP naturally (in vivo) occurring phosphorylation sites and that there was no overlap between the phosphorylation and glycosylation sites. The 5.8 mol P-Ser/mol BSP reflects the total number of mols of naturally occurring phosphorylation, phosphorylated in vivo by CKII (4.1 mol), protein kinase C (0.9 mol), and cGMP-dependent kinase (0.8 mol). Peptide N-terminal sequence analyses of both in vitro (32P) and in vivo (14C) phosphorylated peptides indicated that the phosphorylated residues were predominantly on the N-terminal half of the protein that included recognition sequences for CKII, e.g., LESDEENGVFK (residues 12-22).  相似文献   

2.
The biochemical mechanism controlling nucleation of mineral crystals in developing bone, along with the growth and propagation of these crystals once formed, remains poorly understood. To define the nucleation mechanism, a proteomics analysis was begun on isolated biomineralization foci (BMF), sites of initial crystal nucleation in osteoblastic cell cultures and in primary bone. Comparative analyses of the protein profile for mineralized BMF with that for total osteoblast cultures revealed the latter were enriched in several proteins including BAG-75 and BSP, as well as fragments of each. When 12 protease inhibitors were added separately to UMR 106-01 osteoblastic cultures, only the serine protease inhibitor 4-(2-aminoethyl) benzenesulfonyl fluoride hydrochloride (AEBSF) blocked cleavage of BAG-75 and BSP, and prevented mineral crystal nucleation within BMF. Consideration of the specificities of the inhibitors tested and the fact that AEBSF inhibition was not dependent upon inclusion of FBS in the culture media indicated that mineral nucleation does not require serine protease plasmin, thrombin, kallikrein, urokinase, C1s or furin. In contrast, SKI-1 (S1P or site-1) is a membrane-bound serine protease inhibitable by AEBSF. We show here for the first time that mineralizing UMR 106 cells express a 98-kDa active, soluble form of SKI-1 within BMF. In contrast, nonmineralizing UMR cells appear to differentially process SKI-1 into smaller immunoreactive fragments (<35 kDa). These findings suggest that SKI-1 plays a direct or indirect role in assembly of functional nucleation complexes containing BAG-75 and BSP and their fragments, thus facilitating initial mineral nucleation within these biomineralization foci.  相似文献   

3.
The present study for the first time evaluated both the in vitro and in vivo phosphorylation regions of bone sialoprotein (BSP) by utilizing multiple approaches and techniques. The in vitro phosphorylation sites were determined by 32 P-labeling of native BSP using purified casein kinase II (CKII), followed by peptide mapping and solid-phase N-terminal sequence analyses. The in vivo phosphorylation sites were determined by (i) derivatization with 1-S-[ 14 C]carboxymethyl-dithiothreitol ([ 14 C] CM-DTT) of the proteolytic digests of BSP, followed by isolation and N-terminal peptide sequence analysis; and (ii) analyzing the proteolytic peptides of native BSP using matrix-assisted laser desorption/ionization-time of flight-mass spectrometry (MALDI-TOF-MS). Native BSP incorporated ~2.5 mol of phosphate/mol of BSP by CKII, which were distributed over four major peptide peaks and three shoulder peaks within the peptide map with varying degrees of phosphorylation. Further studies using the [ 14 C] CM-DTT thiol reagent indicated that native and deglycosylated BSP incorporated 5.84 and 5.80 mol of 14 C/mol of BSP, respectively. This confirmed that there were ~5.8 mol P-Ser/mol of BSP naturally (in vivo) occurring phosphorylation sites and that there was no overlap between the phosphorylation and glycosylation sites. The 5.8 mol P-Ser/mol BSP reflects the total number of mols of naturally occurring phosphorylation, phosphorylated in vivo by CKII (4.1 mol), protein kinase C (0.9 mol), and cGMP-dependent kinase (0.8 mol). Peptide N-terminal sequence analyses of both in vitro ( 32 P) and in vivo ( 14 C) phosphorylated peptides indicated that the phosphorylated residues were predominantly on the N-terminal half of the protein that included recognition sequences for CKII, e.g., LESDEENGVFK (residues 12-22).  相似文献   

4.
L Addadi  A Berman  J M Oldak  S Weiner 《Connective tissue research》1989,21(1-4):127-34; discussion 135
Detailed studies of the molecular organization of various mineralized tissues highlight the role played by acidic glycoproteins in controlling crystal formation. In vitro experiments show that these macromolecules are able to interact with specific faces of different crystals, influencing both nucleation and crystal growth. Significantly a common stereochemical motif can be recognized in all the interacting faces studies to date.  相似文献   

5.
《Connective tissue research》2013,54(1-4):127-135
Detailed studies of the molecular organization of various mineralized tissues highlight the role played by acidic glycoproteins in controlling crystal formation. In vitro experiments show that these macromolecules are able to interact with specific faces of different crystals, influencing both nucleation and crystal growth. Significantly a comman stereochemical motif can be recognized in all the interacting faces studies to date.  相似文献   

6.
7.
Crystal deposition in arthritic diseases has attracted much interest. Many reports have established the presence of calcium pyrophosphate (CPPD), hydroxyapatite (HAP) and urate crystals throughout the range of arthritic diseases. In particular, HAP crystals have been detected in 30-60% of synovial fluid (SF) samples from patients suffering from osteoarthritis (OA) and 33% of those suffering from rheumatoid arthritis (RA). In OA, crystal deposition has been linked to greater joint deterioration. The mechanism of intra-articular calcification is unknown. Nucleation is required to transform a 'metastable' phosphate- and calcium-rich biofluid into one that generates crystals. Ferric ions have been demonstrated to induce crystallization of these stable supersaturated solutions via the process of nucleation.The inflamed arthritic joint is prone to iron loading. Microbleeding from compromised vasculature contributes to intra-articular iron loading in arthritic conditions. Low-molecular-mass redox-active iron complexes have been detected in SF in inflammatory joint diseases. These species are credited with mediating oxidative stress via interaction with peroxides and superoxide. In addition, adventitious low-molecular-mass iron complexes can cause nucleation leading to crystal growth within the joint.Decorporating agents capable of removing this misplaced iron from the arthritic joint would have the joint benefit of relieving oxidative stress and preventing crystal nucleation. Systemic side effects could be overcome by the targeting suitable chelators using bioreductive delivery systems that are activated in hypoxic inflamed synovial tissue.  相似文献   

8.
Peptide binding to major histocompatibility complex (MHC) class II molecules is optimal at mildly acidic pH. X-ray crystal structures solved for the murine class II molecule I-E(k) revealed an interesting localization of negatively charged residues within the P6 pocket, which may have implications in the pH dependency of peptide binding. Protonation of these critical residues, under acidic conditions, has been proposed to be important for the formation of stable class II-peptide complexes. In this study, we address a possible role for these charged residues in the pH dependency of peptide binding. An I-E(k) mutant was generated in which two acidic residues of the P6 pocket were substituted with uncharged residues. This class II mutant was expressed, purified, and tested for its ability to bind peptides. The mutant I-E(k) was observed to load peptides optimally at mildly acidic pH. Peptide binding to the mutant was enhanced in the presence of DM, and optimal DM-enhanced binding occurred in the acidic pH range. These findings indicate that structural changes other than protonation of acidic residues in pocket 6 must play a dominant role in pH-regulated peptide binding to I-E(k).  相似文献   

9.
《Connective tissue research》2013,54(1-4):385-392
Bone sialoprotein (BSP) was shown to be a potent nucleator of hydroxy apatite (HA) in a steady-state agarose gel system (Hunter and Goldberg, 1993, PNAS 90: 8562). Nucleation of HA was also demonstrated with the homopolymer poly-glutamic acid but not with poly-aspartic acid or osteopontin. Since BSP contains contiguous sequences of glutamic acid, it is reasonable to suggest that the HA-nucleating activity of BSP resides within these regions. Purified porcine BSP was treated with trypsin and digests fractionated by gel filtration. In addition to small peptides (P3–5), two peptides of 38 kDa (PI) and 25 kDA (P2) were recovered, and after characterization assigned to the regions within BSP encompassing residues 133–272 (PI) and 42–125 (P2). Each of these peptides contained one of the two glutamic acid-rich regions of porcine BSP. In the steady-state agarose gel system, BSP, PI and P2 induced HA formation, whereas the pooled small BSP-derived peptides (P3–5) did not. Analysis by circular dichroism spectroscopy revealed that the homopolymer poly-L-glutamic acid assumes a helical structure, while poly-L-aspartic acid does not. These findings suggest that the nucleating activity does not require intact molecules, that the nucleation of HA by BSP appears to require glutamic acid-rich sequences in a helical conformation and that there are two domains in porcine BSP that are each capable of nucleating HA.  相似文献   

10.
Proteoglycans and their spatial arms, glycosaminoglycans (GAGs), are known to interact with hydroxyapatite (HAP) and have been implicated as important modulators of mineralisation. In the present study isotherm data (0.02 M sodium acetate, pH 6.8) revealed that the iduronic-rich GAGs heparan sulphate, heparin and dermatan sulphate showed greater binding onto HAP with higher adsorption maxima compared with the glucuronic acid-rich GAGs chondroitin-4-sulphate, chondroitin-6-sulphate and hyaluronan. Chemically desulphated chondroitin showed no adsorption onto HAP. With the exception of hyaluronan, the GAGs studied showed no desorbability in sodium acetate buffer only, whereas in di-sodium orthophosphate, desorption occurred much more readily. The data indicates that GAG chemistry and conformation in solution greatly influence the interaction of these molecules with HAP. The conformational flexibility of iduronic acid residues may be an important determinant in the strong binding of iduronic acid-rich GAGs to HAP, increasing the possibility of the appended anionic groups matching calcium sites on the HAP surface, compared with more rigid glucuronic acid residues. This work provides important information concerning interfacial adsorption phenomena between the organic-inorganic phases of mineralised systems.  相似文献   

11.
Porphyromonas gingivalis fimbriae specifically bind salivary acidic proline-rich protein 1 (PRP1) through protein-protein interactions. The binding domains of fimbrillin (a subunit of fimbriae) for PRP1 were analyzed previously (A. Amano, A. Sharma, J.-Y. Lee, H. T. Sojar, P. A. Raj, and R. J. Genco, Infect. Immun. 64:1631-1637, 1996). In this study, we investigated the sites of binding of the PRP1 molecules to the fimbriae. PRP1 (amino acid residues 1 to 150) was proteolysed to three fragments (residues 1 to 74 [fragment 1-74], 75 to 129, and 130 to 150). 125I-labeled fimbriae clearly bound fragments 75-129 and 130-150, immobilized on a polyvinylidene difluoride membrane; both fragments also inhibited whole-cell binding to PRP1-coated hydroxyapatite (HAP) beads by 50 and 83%, respectively. However, the N-terminal fragment failed to show any effect. Analogous peptides corresponding to residues 75 to 89, 90 to 106, 107 to 120, 121 to 129, and 130 to 150 of PRP1 were synthesized. The fimbriae significantly bound peptide 130-150, immobilized on 96-well plates, and the peptide also inhibited binding of 125I-labeled fimbriae to PRP1-coated HAP beads by almost 100%. Peptides 75-89, 90-106, and 121-129, immobilized on plates, showed considerable ability to bind fimbriae. For further analysis of active sites in residues 130 to 150, synthetic peptides corresponding to residues 130 to 137, 138 to 145, and 146 to 150 were prepared. Peptide 138-145 (GRPQGPPQ) inhibited fimbrial binding to PRP1-coated HAP beads by 97%. This amino acid sequence was shared in the alignment of residues 75 to 89, 90 to 106, and 107 to 120. Six synthetic peptides were prepared by serial deletions of individual residues from the N and C termini of peptide GRPQGPPQ. Peptide PQGPPQ was as inhibitory as peptide GRPQGPPQ. Further deletions of the dipeptide Pro-Gln from the N and C termini of peptide PQGPPQ resulted in significant loss of the inhibitory effect. These results strongly suggest that PQGPPQ is the minimal active segment for binding to P. gingivalis fimbriae and that the moiety of the Pro-Gln dipeptide plays a critical role in expressing binding ability.  相似文献   

12.
Vitronectin (VN) is an important matricellular protein that plays a role in cell adhesion, cell migration, wound healing, and inflammation. VN is present in both serum and plasma and studies have shown that it has significant surface enrichment on implanted biomaterials as compared to other proteins present in plasma. Previously, charged self-assembled monolayers (SAMs) have been shown to influence the orientation and conformation of adsorbed proteins and their subsequent bioactivity. In addition, numerous studies have shown that the conformation of VN plays a critical role in its bioactive properties. In this study, both the orientation and conformation of adsorbed VN are systematically studied to determine which of these has the greater influence on its cell binding ability. Atomic force microscopy was used to qualitatively asses the adsorbed amount of VN on positively and negatively charged SAMs both before and after conformation changes induced by pH. Cell adhesion assays were then performed to compare the bovine aortic endothelial cell (BAEC) binding ability of VN in three different conformations on positively and negatively charged surfaces. The results indicate that conformation, not orientation, plays the more important role in the BAEC adhesion properties of VN. In addition, the cell binding properties of VN were maximized when the protein is in the conformation imparted to it by acidic conditions. This indicates that the bioactivity of VN is maximized when in its conformation typical of an acidic wound healing environment, emphasizing the biological relevance of this protein for cell binding under these natural biological conditions.  相似文献   

13.
A novel type of glass-based nanoscale hydroxyapatite (HAP) bioactive bone cement (designed as GBNHAPC) was synthesized by adding nanoscale hydroxyapatite crystalline (20-40 nm), into the self-setting glass-based bone cement (GBC). The inhibition rate of nanoscale HAP and micron HAP on osteosarcoma U2-OS cells was examined. The effects of nanoscale HAP on the crystal phase, microstructure and compressive strength of GBNHAPC were studied, respectively. It was concluded that nanoscale HAP could inhibit the cell proliferation, whereas micron HAP could not, and that nanoscale HAP could be dispersed in the cement evenly and the morphology did not change significantly after a longer immersion time. XRD and FTIR results show nanoscale HAP did not affect the setting reaction of the cement. Furthermore, GBNHAPC had a higher compressive strength (92.6 +/- 3.8 MPa) than GBC (80.1 +/- 3.0 MPa). It was believed that GBNHAPC might be a desirable biomaterial that could not only fill bone defects but also inhibit cancer cell growth.  相似文献   

14.
15.
This article deals with the effect of two acidic polypeptides [polyaspartic acid (PA) and polyglutamic acid (PG)] onto hydroxyapatite (HAP) dissolution by separately considering their influence when they are present only at the HAP interface and when they are both adsorbed and present in the bulk solution. We first determined the amount of adsorbed PA and PG at pH 7.0 and 5.0 onto 10 mg of HAP. Dissolution experiments were performed at pH 5.0 under pH stat conditions by continuously following the consumed protons and released calcium versus time with the aid of specific electrodes. The released phosphate ions were determined by spectrophotometric analysis. The data show that, because of their calcium chelating properties, the polypeptides act as a driving force for HAP dissolution when PA and PG remain present in solution and the interfacial beneficial effect of the adsorbed peptides is erased by the chelating properties of PA and PG present in the solution. When the polypeptides are only adsorbed at the interface, even if a partial PA or PG desorption occurs, HAP dissolution inhibition is still observed.  相似文献   

16.
Fimbriae are considered important in the adherence and colonization of Porphyromonas gingivalis in the oral cavity. It has been demonstrated that purified fimbriae bind to whole human saliva adsorbed to hydroxyapatite (HAP) beads, and the binding appears to be mediated by specific protein-protein interactions. Recently, we expressed the recombinant fimbrillin protein (r-Fim) of P. gingivalis corresponding to amino acid residues 10 to 337 of the native fimbrillin (A. Sharma, H.T. Sojar, J.-Y. Lee, and R.J. Genco, Infect. Immun. 61:3570-3573, 1993). We examined the ability of individual salivary components to promote the direct attachment of r-Fim to HAP beads. Purified r-Fim was radiolabeled with 125I and incubated with HAP beads which were coated with saliva or purified individual salivary components. Whole, parotid, and submandibular-sublingual salivas increased the binding of 125I-r-Fim to HAP beads. Submandibular-sublingual saliva was most effective in increasing the binding of 125I-r-Fim to HAP beads (1.8 times greater than that to uncoated HAP beads). The binding of 125I-r-Fim to HAP beads coated with acidic proline-rich protein 1 (PRP1) or statherin was four and two times greater, respectively, than that to uncoated HAP beads. PRP1 and statherin molecules were also found to bind 125I-r-Fim in an overlay assay. The binding of intact P. gingivalis cells to HAP beads coated with PRP1 or statherin was also enhanced, by 5.4 and 4.3 times, respectively, over that to uncoated HAP beads. The interactions of PRP1 and statherin with 125I-r-Fim were not inhibited by the addition of carbohydrates or amino acids. PRP1 and statherin in solution did not show inhibitory activity on 125I-r-Fim binding to HAP beads coated with PRP1 or statherin. These results suggest that P. gingivalis fimbriae bind strongly through protein-protein interactions to acidic proline-rich protein and statherin molecules which coat surfaces.  相似文献   

17.
Bone sialoprotein (BSP) and bone osteopontin (OPN) are members of the SIBLING (small integrin-binding ligand, N-linked glycoproteins) family of proteins commonly found in mineralized tissues. Previously, OPN was shown to exhibit a preferential orientation for MC3T3-E1 cell adhesion when it was specifically bound to collagen. In this work, the orientation of BSP under similar circumstances is examined and compared with OPN. Radiolabeled adsorption isotherms were obtained for BSP bound to both tissue culture polystyrene (TCPS) and collagen-coated TCPS. The results show that collagen has the capacity to bind almost twice as much OPN under identical conditions. An in vitro MC3T3-E1 cell adhesion assay was then performed to compare the cell binding ability of BSP on either TCPS or collagen-coated TCPS with identical amounts of adsorbed protein. It was found that there is no significant difference in the cell binding ability of BSP on either of the substrates. For cell binding studies on collagen-coated TCPS, it was shown that there are a greater number of cells bound to substrates with adsorbed OPN as compared with BSP. The preferable orientation of OPN for cell binding coupled with the higher binding capability of collagen for OPN indicates that OPN is more important than BSP for osteoblast adhesion to the collagen matrix. In addition, a cell inhibition assay was performed to show that all of the cell binding that occurred throughout these studies was dependent upon integrin interactions with the RGD cell binding moiety.  相似文献   

18.
The phosphorylated acidic glycoproteins bone sialoprotein (BSP) and osteopontin (OPN) bind to hydroxyapatite (HA) crystals and may be involved in the regulation of bone mineralization. The HA-binding properties of these proteins have been attributed to glutamic acid-rich sequences in BSP and aspartic acid-rich sequences in OPN. The present study examines the roles of these polycarboxylate sequences in the binding of BSP and OPN to HA. Porcine BSP, OPN and the synthetic polypeptides poly-L-glutamic acid [Poly(Glu)] and poly-L-aspartic acid [Poly(Asp)] were labeled with fluorescein isothiocyanate and their binding to HA determined by fluorimetry. From the binding isotherms, dissociation constants (KDs) for all the reagents tested were determined to be in the micromolar range. The saturation binding capacities of HA for Poly(Glu), Poly(Asp), BSP and OPN were similar (500-600 micrograms/m2). To investigate the role of glutamic acid-rich and aspartic acid-rich sequences in the binding to HA of BSP and OPN, respectively, competitive binding studies with Poly(Glu) and Poly(Asp) were performed. Poly(Glu) was able to displace a maximum of 100% of Poly(Glu), 81% of OPN, 68% of BSP and 65% of Poly(Asp). Poly(Asp) was able to displace a maximum of 100% of Poly(Glu), 99% of Poly(Asp), 95% of OPN and 89% of BSP. These results are consistent with the view that BSP and OPN bind to HA via their polycarboxylate sequences, but suggest a complex mode of interaction between polyelectrolytes and ionic crystals.  相似文献   

19.
HAP1 positive control mutants specific for one of two binding sites.   总被引:8,自引:0,他引:8  
  相似文献   

20.
Integrin activation and structural rearrangement   总被引:26,自引:0,他引:26  
Summary: Among adhesion receptor families, integrins are particularly important in biological processes that require rapid modulation of adhesion and de-adhesion. Activation on a timescale of < 1 s of β2 integrins on leukocytes and β3 integrins on platelets enables deposition of these cells at sites of inflammation or vessel wall injury. Recent crystal, nuclear magnetic resonance (NMR), and electron microscope (EM) structures of integrins and their domains lead to a unifying mechanism of activation for both integrins that contain and those that lack an inserted (I) domain. The I domain adopts two alternative conformations, termed open and closed. In striking similarity to signaling G-proteins, rearrangement of a Mg2+-binding site is linked to large conformational movements in distant backbone regions. Mutations that stabilize a particular conformation show that the open conformation has high affinity for ligand, whereas the closed conformation has low affinity. Movement of the C-terminal α-helix 10 Å down the side of the domain in the open conformation is sufficient to increase affinity at the distal ligand-binding site 9000-fold. This C-terminal "bell-rope" provides a mechanism for linkage to conformational movements in other domains. Recent structures and functional studies reveal interactions between β-propeller, I, and I-like domains in the integrin headpiece, and a critical role for integrin epidermal growth factor (EGF) domains in the stalk region. The headpiece of the integrin faces down towards the membrane in the inactive conformation, and extends upward in a "switchblade"-like opening upon activation. These long-range structural rearrangements of the entire integrin molecule involving interdomain contacts appear closely linked to conformational changes within the I and I-like domains, which result in increased affinity and competence for ligand binding.  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