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.
As a hemoprotein, hemoglobin (Hb) can, in the presence of H2O2, act as a peroxidase. In red blood cells, this activity is regulated by the reducing environment. For stroma-free Hb this regulation is lost, and the potential for Hb to become a peroxidase is high and further increased by inflammatory cells generating superoxide. The latter can be converted into H2O2 and feed Hb peroxidase activity. Haptoglobins (Hp) bind with extracellular Hb and reportedly weaken Hb peroxidase activity. Here we demonstrate that: (i) Hb peroxidase activity is retained upon binding with Hp; (ii) in the presence of H2O2, Hb·Hp peroxidase complexes undergo covalent cross-linking; (iii) peroxidase activity of Hb·Hp complexes and aggregates consumes reductants such as ascorbate and nitric oxide; (iv) cross-linked Hb·Hp aggregates are taken up by macrophages at rates exceeding those for noncovalently cross-linked Hb·Hp complexes; (v) the engulfed Hb·Hp aggregates activate superoxide production and induce intracellular oxidative stress (deplete endogenous glutathione and stimulate lipid peroxidation); (vi) Hb·Hp aggregates cause cytotoxicity to macrophages; and (vii) Hb·Hp aggregates are present in septic plasma. Overall, our data suggest that under conditions of severe inflammation and oxidative stress, peroxidase activity of Hb·Hp covalent aggregates may cause macrophage dysfunction and microvascular vasoconstriction, which are commonly seen in severe sepsis and hemolytic diseases.As a hemoprotein, Hb,2 in the presence of oxidizing equivalents such as H2O2, can act as a peroxidase with very high oxidizing potential (1). In red blood cells, this potentially dangerous activity is strictly regulated by the reducing environment and the lack of oxidizing equivalents. The inadvertently appearing ferric forms of Hb are short-lived, and the hemoprotein is effectively converted into ferro-Hb (deoxy-Hb) by metHb reductase (2). Normally, less than 2% of total Hb exists in the form of MetHb because the rate of Hb reduction is far greater than its oxidation (2). For stroma-free Hb, however, this intracellular regulation is lost, and the likelihood for Hb to act as a peroxidase is high. This possibility is markedly increased in the course of severe inflammation (e.g. in sepsis) by the generation of superoxide radicals by immune cells. The latter can be spontaneously or catalytically (by extracellular superoxide dismutase) converted into H2O2, a fuel for Hb peroxidase activity. In line with this, several clinical and experimental investigations have established that lethality in sepsis is increased in the setting of hemolysis (35).Circulating haptoglobin (Hp) provides an important endogenous defense against the toxic effects of Hb (6, 7). The major biological function of this abundant plasma protein is binding and recycling of stroma-free Hb via the macrophage CD163 receptor-mediated pathway (8, 9). It has been proposed that Hp possesses antioxidant activity and diminishes oxidative stress induced by stroma-free Hb (1012). The antioxidant action of Hp toward Hb has been associated, at least in part, with weakening of its peroxidase activity (10) or preventing oxidation and cross-linking of Hb (7).Previous work has demonstrated that peroxidase activity of different hemoproteins, including Hb, can induce protein self-oxidation leading to covalent cross-linking and aggregation (1316). Whether these hetero-oligomeric covalent aggregates retain the peroxidase activity is unknown. If the aggregates retain peroxidase activity, they may continue to be a source of oxidative stress both in circulation as well as in phagocytizing cells involved in their clearance such as macrophages.In the current work, we determined the extent to which: (i) Hb peroxidase activity is decreased by binding with Hp; (ii) peroxidase activity of Hb·Hp complexes initiates cross-linking into covalent hetero-oligomers; (iii) peroxidase activity of Hb·Hp complexes and aggregates utilizes nitric oxide (NO); (iv) Hb·Hp aggregates are taken up by macrophages as compared with noncovalent Hb·Hp complexes; and (v) the engulfed Hb·Hp aggregates induce oxidative stress and cytotoxicity. Here, we report that Hb·Hp complexes and aggregates are potent peroxidases capable of inducing oxidative stress in both plasma and macrophages. We further demonstrate the presence of Hb·Hp aggregates in septic plasma.  相似文献   

2.
α-Hemoglobin (αHb) stabilizing protein (AHSP) is expressed in erythropoietic tissues as an accessory factor in hemoglobin synthesis. AHSP forms a specific complex with αHb and suppresses the heme-catalyzed evolution of reactive oxygen species by converting αHb to a conformation in which the heme is coordinated at both axial positions by histidine side chains (bis-histidyl coordination). Currently, the detailed mechanism by which AHSP induces structural changes in αHb has not been determined. Here, we present x-ray crystallography, NMR spectroscopy, and mutagenesis data that identify, for the first time, the importance of an evolutionarily conserved proline, Pro30, in loop 1 of AHSP. Mutation of Pro30 to a variety of residue types results in reduced ability to convert αHb. In complex with αHb, AHSP Pro30 adopts a cis-peptidyl conformation and makes contact with the N terminus of helix G in αHb. Mutations that stabilize the cis-peptidyl conformation of free AHSP, also enhance the αHb conversion activity. These findings suggest that AHSP loop 1 can transmit structural changes to the heme pocket of αHb, and, more generally, highlight the importance of cis-peptidyl prolyl residues in defining the conformation of regulatory protein loops.Mammalian adult hemoglobin (HbA)5 is a tetramer of two αHb and two βHb subunits, which is produced to extremely high concentrations (∼340 mg/ml) in red blood cells. Numerous mechanisms exist to balance and coordinate HbA synthesis in normal erythropoiesis, and problems with the production of either HbA subunit give rise to thalassemia, a common cause of anemia worldwide. Previously, we identified α-hemoglobin stabilizing protein (AHSP) as an accessory factor in normal HbA production (1). AHSP forms a dimeric complex with αHb (see Fig. 1A) (2) but does not interact with βHb or HbA. AHSP also binds heme-free (apo) αHb (3) and may serve functions in both the folding of nascent αHb (4) and the detoxification of excess αHb that remains following HbA assembly (2, 5). Mice carrying an Ahsp gene knock-out display mild anemia, ineffective erythropoiesis, and enhanced sensitivity to oxidative stress (1, 6), features also observed in β-thalassemia patients due to the cytotoxic effects of free αHb.Open in a separate windowFIGURE 1.Summary of αHb·AHSP interactions. A, the αHb·AHSP complex(PDB code 1Z8U) (2). The interface is formed from helices 1 and 2 and the intervening loop 1 (green) of AHSP, together with helices G-H and the B-C corner of αHb (cyan). B, detailed views of the heme binding site of αHb as it appears in oxy-HbA (PDB code 1GZX) (69) and the final bis-histidyl αHb·AHSP complex (PDB code 1Z8U) with two histidine ligands to the iron. Typical visible absorption spectra in the region 450–700 nm are shown.Free αHb promotes the formation of harmful reactive oxygen species as a result of reduction/oxidation reactions involving the heme iron (7, 8). Reactive oxygen species can damage heme, αHb, and other cellular structures, resulting in hemoglobin precipitates and death of erythroid precursor cells (912). The presence of AHSP may explain how cells tolerate the slight excess of αHb that is observed in normal erythropoiesis, which is postulated to inhibit the formation of non-functional βHb tetramers, thus providing a robust mechanism for achieving the correct subunit stoichiometry during HbA assembly (13).Structural and biochemical studies have begun to elucidate the molecular mechanism by which AHSP detoxifies αHb. AHSP binds to oxygenated αHb to generate an initial complex that retains the oxy-heme, as evidenced by a characteristic visible absorption spectrum (see Fig. 1B, middle) and resonance Raman spectrum (5). This initial oxy-αHb·AHSP complex then converts to a low spin Fe3+ complex (2), in which the heme iron is bound at both axial positions by the side chains of His58 and His87 from αHb (see Fig. 1B, right). The formation of this complex inhibits αHb peroxidase activity and heme loss (2). Bis-histidyl heme coordination is becoming increasingly recognized as a feature of numerous vertebrate and non-vertebrate globins (14) and has been shown previously to confer a relative stabilization of the Fe3+ over the Fe2+ oxidation state (1517). Although bis-histidyl heme coordination has previously been detected in solutions of met-Hb, formed through spontaneous autoxidation of Hb (1821), the bishis-αHb·AHSP complex provides the first evidence that the bis-histidyl heme may play a positive functional role in Hb biochemistry by inhibiting the production of harmful reactive oxygen species.Despite its potential importance, the mechanism by which AHSP influences heme coordination in its binding partner is still unknown. As shown in Fig. 1A, AHSP binds αHb at a surface away from the heme pocket, and thus structural changes must somehow be transmitted through the αHb protein. It is intriguing that the free AHSP protein switches between two alternative conformations linked to cis/trans isomerization of the Asp29-Pro30 peptide bond in loop 1 (22) and that, in complex with αHb, this loop is located at the αHb·AHSP interface (see Fig. 1A). Peptide bonds preceding proline residues are unique in that the cis or trans bonding conformations have relatively similar stabilities (23), allowing an interconversion between these conformations that can be important for protein function (24, 25). Previous x-ray crystal structures of αHb·AHSP complexes have been obtained only with a P30A mutant of AHSP, in which isomerization is abolished and the Asp29-Ala30 peptide bond adopts a trans conformation, leaving the potential structural and functional significance of the evolutionarily conserved Pro30 undisclosed. Here, we demonstrate a functional role for AHSP Pro30 in conversion of oxy-αHb to the bis-histidyl form and identify a specific structural role for a cis Asp29-Pro30 peptide bond in this process. From a mechanistic understanding of how AHSP promotes formation of bis-histidyl αHb, we may eventually be able to engineer AHSP function as a tool in new treatments for Hb diseases such as β-thalassemia.  相似文献   

3.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

4.
Clinically, amniotic membrane (AM) suppresses inflammation, scarring, and angiogenesis. AM contains abundant hyaluronan (HA) but its function in exerting these therapeutic actions remains unclear. Herein, AM was extracted sequentially with buffers A, B, and C, or separately by phosphate-buffered saline (PBS) alone. Agarose gel electrophoresis showed that high molecular weight (HMW) HA (an average of ∼3000 kDa) was predominantly extracted in isotonic Extract A (70.1 ± 6.0%) and PBS (37.7 ± 3.2%). Western blot analysis of these extracts with hyaluronidase digestion or NaOH treatment revealed that HMW HA was covalently linked with the heavy chains (HCs) of inter-α-inhibitor (IαI) via a NaOH-sensitive bond, likely transferred by the tumor necrosis factor-α stimulated gene-6 protein (TSG-6). This HC·HA complex (nHC·HA) could be purified from Extract PBS by two rounds of CsCl/guanidine HCl ultracentrifugation as well as in vitro reconstituted (rcHC·HA) by mixing HMW HA, serum IαI, and recombinant TSG-6. Consistent with previous reports, Extract PBS suppressed transforming growth factor-β1 promoter activation in corneal fibroblasts and induced mac ro phage apo pto sis. However, these effects were abolished by hyaluronidase digestion or heat treatment. More importantly, the effects were retained in the nHC·HA or rcHC·HA. These data collectively suggest that the HC·HA complex is the active component in AM responsible in part for clinically observed anti-inflammatory and anti-scarring actions.Hyaluronan (HA)4 is widely distributed in extracellular matrices, tissues, body fluids, and even in intracellular compartments (reviewed in Refs. 1 and 2). The molecular weight of HA ranges from 200 to 10,000 kDa depending on the source (3), but can also exist as smaller fragments and oligosaccharides under certain physiological or pathological conditions (1). Investigations over the last 15 years have suggested that low Mr HA can induce the gene expression of proinflammatory mediators and proangiogenesis, whereas high molecular weight (HMW) HA inhibits these processes (47).Several proteins have been shown to bind to HA (8) such as aggrecan (9), cartilage link protein (10), versican (11), CD44 (12, 13), inter-α-inhibitor (IαI) (14, 15), and tumor necrosis factor-α stimulated gene-6 protein (TSG-6) (16, 17). IαI consists of two heavy chains (HCs) (HC1 and HC2), both of which are linked through ester bonds to a chondroitin sulfate chain that is attached to the light chain, i.e. bikunin. Among all HA-binding proteins, only the HCs of IαI have been clearly demonstrated to be covalently coupled to HA (14, 18). However, TSG-6 has also been reported to form stable, possibly covalent, complexes with HA, either alone (19, 20) or when associated with HC (21).The formation of covalent bonds between HCs and HA is mediated by TSG-6 (2224) where its expression is often induced by inflammatory mediators such as tumor necrosis factor-α and interleukin-1 (25, 26). TSG-6 is also expressed in inflammatory-like processes, such as ovulation (21, 27, 28) and cervical ripening (29). TSG-6 interacts with both HA (17) and IαI (21, 24, 3033), and is essential for covalently transferring HCs on to HA (2224). The TSG-6-mediated formation of the HC·HA complex has been demonstrated to play a crucial role in female fertility in mice. The HC·HA complex is an integral part of an expanded extracellular “cumulus” matrix around the oocyte, which plays a critical role in successful ovulation and fertilization in vivo (22, 34). HC·HA complexes have also been found at sites of inflammation (3538) where its pro- or anti-inflammatory role remain arguable (39, 40).Immunostaining reveals abundant HA in the avascular stromal matrix of the AM (41, 42).5 In ophthalmology, cryopreserved AM has been widely used as a surgical graft for ocular surface reconstruction and exerts clinically observable actions to promote epithelial wound healing and to suppress inflammation, scarring, and angiogenesis (for reviews see Refs. 4345). However, it is not clear whether HA in AM forms HC·HA complex, and if so whether such an HC·HA complex exerts any of the above therapeutic actions. To address these questions, we extracted AM with buffers of increasing salt concentration. Because HMW HA was found to form the HC·HA complex and was mainly extractable by isotonic solutions, we further purified it from the isotonic AM extract and reconstituted it in vitro from three defined components, i.e. HMW HA, serum IαI, and recombinant TSG-6. Our results showed that the HC·HA complex is an active component in AM responsible for the suppression of TGF-β1 promoter activity, linkable to the scarring process noted before by AM (4648) and by the AM soluble extract (49), as well as for the promotion of macrophage death, linkable to the inflammatory process noted by AM (50) and the AM soluble extract (51).  相似文献   

5.
6.
7.
Although PTIP is implicated in the DNA damage response, through interactions with 53BP1, the function of PTIP in the DNA damage response remain elusive. Here, we show that RNF8 controls DNA damage-induced nuclear foci formation of PTIP, which in turn regulates 53BP1 localization to the DNA damage sites. In addition, SMC1, a substrate of ATM, could not be phosphorylated at the DNA damage sites in the absence of PTIP. The PTIP-dependent pathway is important for DNA double strand breaks repair and DNA damage-induced intra-S phase checkpoint activation. Taken together, these results suggest that the role of PTIP in the DNA damage response is downstream of RNF8 and upstream of 53BP1. Thus, PTIP regulates 53BP1-dependent signaling pathway following DNA damage.The DNA damage response pathways are signal transduction pathways with DNA damage sensors, mediators, and effectors, which are essential for maintaining genomic stability (13). Following DNA double strand breaks, histone H2AX at the DNA damage sites is rapidly phosphorylated by ATM/ATR/DNAPK (410), a family homologous to phosphoinositide 3-kinases (11, 12). Subsequently, phospho-H2AX (γH2AX) provides the platform for accumulation of a larger group of DNA damage response factors, such as MDC1, BRCA1, 53BP1, and the MRE11·RAD50·NBS1 complex (13, 14), at the DNA damage sites. Translocalization of these proteins to the DNA double strand breaks (DSBs)3 facilitates DNA damage checkpoint activation and enhances the efficiency of DNA damage repair (14, 15).Recently, PTIP (Pax2 transactivation domain-interacting protein, or Paxip) has been identified as a DNA damage response protein and is required for cell survival when exposed to ionizing radiation (IR) (1, 1618). PTIP is a 1069-amino acid nuclear protein and has been originally identified in a yeast two-hybrid screening as a partner of Pax2 (19). Genetic deletion of the PTIP gene in mice leads to early embryonic lethality at embryonic day 8.5, suggesting that PTIP is essential for early embryonic development (20). Structurally, PTIP contains six tandem BRCT (BRCA1 carboxyl-terminal) domains (1618, 21). The BRCT domain is a phospho-group binding domain that mediates protein-protein interactions (17, 22, 23). Interestingly, the BRCT domain has been found in a large number of proteins involved in the cellular response to DNA damages, such as BRCA1, MDC1, and 53BP1 (7, 2429). Like other BRCT domain-containing proteins, upon exposure to IR, PTIP forms nuclear foci at the DSBs, which is dependent on its BRCT domains (1618). By protein affinity purification, PTIP has been found in two large complexes. One includes the histone H3K4 methyltransferase ALR and its associated cofactors, the other contains DNA damage response proteins, including 53BP1 and SMC1 (30, 31). Further experiments have revealed that DNA damage enhances the interaction between PTIP and 53BP1 (18, 31).To elucidate the DNA damage response pathways, we have examined the upstream and downstream partners of PTIP. Here, we report that PTIP is downstream of RNF8 and upstream of 53BP1 in response to DNA damage. Moreover, PTIP and 53BP1 are required for the phospho-ATM association with the chromatin, which phosphorylates SMC1 at the DSBs. This PTIP-dependent pathway is involved in DSBs repair.  相似文献   

8.
Redox-active copper is implicated in the pathogenesis of Alzheimer disease (AD), β-amyloid peptide (Aβ) aggregation, and amyloid formation. Aβ·copper complexes have been identified in AD and catalytically oxidize cholesterol and lipid to generate H2O2 and lipid peroxides. The site and mechanism of this abnormality is not known. Growing evidence suggests that amyloidogenic processing of the β-amyloid precursor protein (APP) occurs in lipid rafts, membrane microdomains enriched in cholesterol. β- and γ-secretases, and Aβ have been identified in lipid rafts in cultured cells, human and rodent brains, but the role of copper in lipid raft amyloidogenic processing is presently unknown. In this study, we found that copper modulates flotillin-2 association with cholesterol-rich lipid raft domains, and consequently Aβ synthesis is attenuated via copper-mediated inhibition of APP endocytosis. We also found that total cellular copper is associated inversely with lipid raft copper levels, so that under intracellular copper deficiency conditions, Aβ·copper complexes are more likely to form. This explains the paradoxical hypermetallation of Aβ with copper under tissue copper deficiency conditions in AD.Imbalance of metal ions has been recognized as one of the key factors in the pathogenesis of Alzheimer disease (AD).2 Aberrant interactions between copper or zinc with the β-amyloid peptide (Aβ) released into the glutamatergic synaptic cleft vicinity could result in the formation of toxic Aβ oligomers and aggregation into plaques characteristic of AD brains (reviewed in Ref. 1). Copper, iron, and zinc are highly concentrated in extracellular plaques (2, 3), and yet brain tissues from AD (46) and human β-amyloid precursor protein (APP) transgenic mice (710) are paradoxically copper deficient compared with age-matched controls. Elevation of intracellular copper levels by genetic, dietary, and pharmacological manipulations in both AD transgenic animal and cell culture models is able to attenuate Aβ production (7, 9, 1115). However, the underlying mechanism is at present unclear.Abnormal cholesterol metabolism is also a contributing factor in the pathogenesis of AD. Hypercholesterolemia increases the risk of developing AD-like pathology in a transgenic mouse model (16). Epidemiological and animal model studies show that a hypercholesterolemic diet is associated with Aβ accumulation and accelerated cognitive decline, both of which are further aggravated by high dietary copper (17, 18). In contrast, biochemical depletion of cholesterol using statins, inhibitors of 3-hydroxy-3-methyglutaryl coenzyme A reductase, and methyl-β-cyclodextrin, a cholesterol sequestering agent, inhibit Aβ production in animal and cell culture models (1925).Cholesterol is enriched in lipid rafts, membrane microdomains implicated in Aβ generation from APP cleavage by β- and γ-secretases. Recruitment of BACE1 (β-secretase) into lipid rafts increases the production of sAPPβ and Aβ (23, 26). The β-secretase-cleaved APP C-terminal fragment (β-CTF), and γ-secretase, a multiprotein complex composed of presenilin (PS1 or PS2), nicastrin (Nct), PEN-2 and APH-1, colocalize to lipid rafts (27). The accumulation of Aβ in lipid rafts isolated from AD and APP transgenic mice brains (28) provided further evidence that cholesterol plays a role in APP processing and Aβ generation.Currently, copper and cholesterol have been reported to modulate APP processing independently. However, evidence indicates that, despite tissue copper deficiency, Aβ·Cu2+ complexes form in AD that catalytically oxidize cholesterol and lipid to generate H2O2 and lipid peroxides (e.g. hydroxynonenal and malondialdehyde), which contribute to oxidative damage observed in AD (2935). The underlying mechanism leading to the formation of pathological Aβ·Cu2+ complexes is unknown. In this study, we show that copper alters the structure of lipid rafts, and attenuates Aβ synthesis in lipid rafts by inhibition of APP endocytosis. We also identify a paradoxical inverse relationship between total cellular copper levels and copper distribution to lipid rafts, which appear to possess a privileged pool of copper where Aβ is more likely to interact with Cu2+ under copper-deficiency conditions to form Aβ·Cu2+ complexes. These data provide a novel mechanism by which cellular copper deficiency in AD could foster an environment for potentially adverse interactions between Aβ, copper, and cholesterol in lipid rafts.  相似文献   

9.
Streptokinase (SK) conformationally activates the central zymogen of the fibrinolytic system, plasminogen (Pg). The SK·Pg* catalytic complex binds Pg as a specific substrate and cleaves it into plasmin (Pm), which binds SK to form the SK·Pm complex that propagates Pm generation. Catalytic complex formation is dependent on lysine-binding site (LBS) interactions between a Pg/Pm kringle and the SK COOH-terminal Lys414. Pg substrate recognition is also LBS-dependent, but the kringle and SK structural element(s) responsible have not been identified. SK mutants lacking Lys414 with Ala substitutions of charged residues in the SK β-domain 250-loop were evaluated in kinetic studies that resolved conformational and proteolytic Pg activation. Activation of [Lys]Pg and mini-Pg (containing only kringle 5 of Pg) by SK with Ala substitutions of Arg253, Lys256, and Lys257 showed decreases in the bimolecular rate constant for Pm generation, with nearly total inhibition for the SK Lys256/Lys257 double mutant. Binding of bovine Pg (BPg) to the SK·Pm complex containing fluorescently labeled Pm demonstrated LBS-dependent assembly of a SK·labeled Pm·BPg ternary complex, whereas BPg did not bind to the complex containing the SK Lys256/Lys257 mutant. BPg was activated by SK·Pm with a Km indistinguishable from the KD for BPg binding to form the ternary complex, whereas the SK Lys256/Lys257 mutant did not support BPg activation. We conclude that SK residues Arg253, Lys256, and Lys257 mediate Pg substrate recognition through kringle 5 of the [Lys]Pg and mini-Pg substrates. A molecular model of the SK·kringle 5 complex identifies the putative interactions involved in LBS-dependent Pg substrate recognition.Streptokinase (SK)6 activates the human fibrinolytic system by activating plasminogen (Pg) through a unique mechanism that is responsible for the use of SK as a thrombolytic drug and its role as a key pathogenicity factor in Group A streptococcal infection (1, 2). The crystal structure of SK bound to the catalytic domain of plasmin (μPm) shows that SK consists of three β-grasp, tightly folded domains, α, β, and γ, linked by flexible segments (3). In solution, SK is highly flexible and behaves hydrodynamically like three beads on a string (4). When bound to μPm, SK assumes a highly ordered structure resembling a three-sided crater surrounding the catalytic site that provides an exosite(s) for binding the catalytic domain of Pg as a substrate (3, 5). In the first step of the SK-mediated Pg activation pathway, SK binds the catalytic domain of the Pg zymogen in a rapid equilibrium process and inserts its NH2-terminal Ile1 residue into the NH2-terminal binding cleft of Pg, activating the catalytic site nonproteolytically (610). Although structural proof is lacking, SK Ile1 presumably forms a critical salt bridge with Asp740(194) (plasminogen numbering; chymotrypsinogen numbering is in parentheses) that initiates conformational activation of the substrate binding site and oxyanion hole required for proteolytic activity (6, 810). The activated SK·Pg* complex binds a second molecule of Pg as a specific substrate and cleaves it at Arg561(15)-Val562(16) to form the fibrin-degrading proteinase, plasmin (Pm) (1014). Proteolytic generation of Pm is propagated by formation of a high affinity SK·Pm complex that converts the remaining free Pg into Pm (5, 11).[Glu]Pg, the full-length form of Pg circulating in blood, consists of an NH2-terminal PAN (Pg/Apple/Nematode (15, 16)) module, followed by five kringle domains (K1–K5), and the trypsin-like serine proteinase catalytic domain (17). Formation of the SK·Pg* and SK·Pm catalytic complexes and Pg substrate binding are inhibited by the lysine analog, 6-aminohexanoic acid (6-AHA), which binds to lysine-binding sites (LBS) located primarily in kringles K1, K4, and K5 of Pg and Pm (10, 11, 1823). Cleavage of the Lys77-Lys78 peptide bond in [Glu]Pg by Pm releases the PAN module and generates the truncated form, [Lys]Pg. Formation of [Lys]Pg is accompanied by a conformational change of [Glu]Pg from a compact, closed α-conformation to a partially extended β-conformation with expression of higher affinity LBS for 6-AHA (24, 25). The fourth kringle module mediates a second conformational change, from the β-conformation to the extended γ-conformation (25).Binding of SK to [Glu]Pg is independent of LBS, with a dissociation constant of 100–150 nm, whereas formation of SK·[Lys]Pg is LBS-dependent with a 13–20-fold higher affinity that is reduced to that of [Glu]Pg by saturating concentrations of 6-AHA (10, 21). Activation of the catalytic domain in [Lys]Pm increases affinity for SK about 830-fold, which is reduced 11–20-fold by 6-AHA (5, 21). Interaction of the COOH-terminal Lys414 residue of SK with a Pg/Pm kringle domain is responsible for the LBS-dependent enhancement of the affinity of SK·[Lys]Pg* and SK·Pm catalytic complex formation (22). Recent rapid reaction kinetic studies of the SK·Pm binding pathway demonstrated that interaction of Lys414 with a Pm kringle enhances formation of an initial rapid equilibrium SK·Pm encounter complex, succeeded by two sequential, tightening conformational changes, to achieve an overall dissociation constant of ∼12 pm (26). The Pg/Pm kringle domain responsible for the enhancement of SK·Pg* and SK·Pm complex formation is not known. Productive interaction of Pg as a substrate of the SK·Pg*/Pm complexes is also greatly inhibited by saturating 6-AHA (11). Kinetic and equilibrium binding studies of SK-mediated Pm formation resolved the conformational activation process from the coupled proteolytic generation of Pm (10, 11). The kinetic approach demonstrated that Lys414 deletion reduced the affinity of formation of the SK·Pg* catalytic complex specifically, whereas the subsequent LBS-dependent proteolytic formation of Pm was unaffected, indicating that Pg substrate recognition is mediated by a structurally distinct region of SK and an unknown kringle (22).Previous structure-function studies have yielded diverse interpretations and conclusions regarding the structural basis of LBS-dependent Pg substrate recognition (23, 2734). Each of the three domains of SK has been implicated in this regard (29, 30, 35, 36), and binding of two Pg molecules to the residue 1–59 sequence of the α-domain has been reported (36). In particular, segments 16–36, 41–48, 48–59, and 88–97 of the SK α-domain have been concluded to play a role in Pg substrate recognition (32, 33, 37, 38). For several SK mutants, a complex mixture of functional effects on their binding to [Glu]Pg and its conformational and proteolytic activation has been reported (28, 31, 33). Some of these effects may result from the inherent flexibility of SK when bound to Pg or Pm (39), and others may be due to the use of kinetic approaches that do not clearly discriminate between conformational and proteolytic activation.Some observations implicate a protruding hairpin loop called the 250-loop (residues Ala251–Ile264) in the SK β-domain in Pg substrate recognition (27, 28, 31, 34). This loop is disordered in the structure of the SK·μPm complex but is ordered in the structure of the isolated β-domain (3, 40). Deletion of the 250-loop, Ala substitution of Lys256 and Lys257 at the apex of the loop, and substitution of multiple residues near and within the loop resulted in disparate effects on Km and kcat for [Glu]Pg activation (27, 28, 31). The conclusions of these studies were that Lys256 and Lys257 are involved in SK binding and conformational activation of [Glu]Pg in addition to proteolytic processing of Pg as a substrate. Some of these studies are problematic because the natural NH2-terminal Ile1 residue necessary for conformational activation is preceded either by an additional methionine (27, 31) or maltose-binding protein (28) in the recombinant SK species used.Because of the diverse conclusions regarding the functional properties of the 250-loop mutations and the possibility of other potential Pg substrate binding sites, the present studies were undertaken to resolve the function of residues in the 250-loop in LBS-dependent Pg substrate recognition by the SK·Pg* complex. The kringle domain of Pg involved in Pg substrate recognition has not been clearly identified but has been suggested to be K5 (27) on the basis that the isolated β-domain bound Pg (30) and K5 (29) in an LBS-dependent manner. Given the general specificity of Pg kringles for COOH-terminal Lys residues and zwitterionic ligands, such as 6-AHA, and the internal sequence of the 250-loop, it appeared possible that a pseudolysine motif on SK was involved. In the binding of a 30-residue peptide from plasminogen binding Group A streptococcal M-like protein (PAM), VEK-30, to K2 of Pg, Castellino and co-workers (41, 42) showed by crystallography and mutagenesis that residues with cationic (Arg and His) and anionic side chains (Glu) arranged spatially on a helix constituted a pseudolysine structure similar to 6-AHA that binds specifically to the LBS of K2. Additional evidence for pseudolysine structures in Pg binding comes from studies of α-enolase from Streptococcus pneumoniae, which has a 9-residue internal binding site for Pg containing essential basic (two Lys residues) and acidic (Asp and Glu residues) located on a surface loop (43, 44).To determine whether a similar SK structure is involved in [Lys]Pg substrate recognition, anionic and cationic residues in the 250-loop were substituted with Ala and characterized in kinetic studies using methods that resolve conformational and proteolytic activation. Studies with [Lys]Pg and mini-Pg, which contains only K5 and the catalytic domain, showed that Arg253, Lys256, and Lys257 facilitate LBS-dependent substrate recognition through interactions with K5. The absence of evidence for a pseudolysine structure in the 250-loop is compatible with the established atypical specificity of K5 for cationic ligands, such as benzamidine, Nα-acetyl-Lys-methyl ester, 6-aminohexane, and 5-aminopentane, in addition to zwitterionic ligands (19, 4547). The studies resolve for the first time the structural features of SK that mediate the LBS-dependent interactions that enhance affinity of SK·Pg* and SK·Pm catalytic complex formation and those that facilitate binding of Pg as a substrate of these complexes.  相似文献   

10.
We have previously reported that growth factor receptor-bound protein-7 (Grb7), an Src-homology 2 (SH2)-containing adaptor protein, enables interaction with focal adhesion kinase (FAK) to regulate cell migration in response to integrin activation. To further elucidate the signaling events mediated by FAK·Grb7 complexes in promoting cell migration and other cellular functions, we firstly examined the phos pho ryl a ted tyrosine site(s) of Grb7 by FAK using an in vivo mutagenesis. We found that FAK was capable of phos pho rylating at least 2 of 12 tyrosine residues within Grb7, Tyr-188 and Tyr-338. Moreover, mutations converting the identified Tyr to Phe inhibited integrin-dependent cell migration as well as impaired cell proliferation but not survival compared with the wild-type control. Interestingly, the above inhibitory effects caused by the tyrosine phos pho ryl a tion-deficient mutants are probably attributed to their down-regulation of phospho-Tyr-397 of FAK, thereby implying a mechanism by competing with wild-type Grb7 for binding to FAK. Consequently, these tyrosine phos pho ryl a tion-deficient mutants evidently altered the phospho-Tyr-118 of paxillin and phos pho ryl a tion of ERK1/2 but less on phospho-Ser-473 of AKT, implying their involvement in the FAK·Grb7-mediated cellular functions. Additionally, we also illustrated that the formation of FAK·Grb7 complexes and Grb7 phos pho ryl a tion by FAK in an integrin-dependent manner were essential for cell migration, proliferation and anchorage-independent growth in A431 epidermal carcinoma cells, indicating the importance of FAK·Grb7 complexes in tumorigenesis. Our data provide a better understanding on the signal transduction event for FAK·Grb7-mediated cellular functions as well as to shed light on a potential therapeutic in cancers.Growth factor receptor bound protein-7 (Grb7)2 is initially identified as a SH2 domain-containing adaptor protein bound to the activated EGF receptor (1). Grb7 is composed of an N-terminal proline-rich region, following a putative RA (Ras-associating) domain and a central PH (pleckstrin homology) domain and a BPS motif (between PH and SH2 domains), and a C-terminal SH2 domain (26). Despite the lack of enzymatic activity, the presence of multiple protein-protein interaction domains allows Grb7 family adaptor proteins to participate in versatile signal transduction pathways and, therefore, to regulate many cellular functions (46). A number of signaling molecules has been reported to interact with these featured domains, although most of the identified Grb7 binding partners are mediated through its SH2 domain. For example, the SH2 domain of Grb7 has been demonstrated to be capable of binding to the phospho-tyrosine sites of EGF receptor (1), ErbB2 (7), ErbB3 and ErbB4 (8), Ret (9), platelet-derived growth factor receptor (10), insulin receptor (11), SHPTP2 (12), Tek/Tie2 (13), caveolin (14), c-Kit (15), EphB1 (16), G6f immunoreceptor protein (17), Rnd1 (18), Shc (7), FAK (19), and so on. The proceeding α-helix of the PH domain of Grb7 is the calmodulin-binding domain responsible for recruiting Grb7 to plasma membrane in a Ca2+-dependent manner (20), and the association between the PH domain of Grb7 and phosphoinositides is required for the phosphorylation by FAK (21). Two additional proteins, NIK (nuclear factor κB-inducing kinase) and FHL2 (four and half lim domains isoform 2), in association with the GM region (Grb and Mig homology region) of Grb7 are also reported, although the physiological functions for these interactions remain unknown (22, 23). Recently, other novel roles in translational controls and stress responses through the N terminus of Grb7 are implicated for the findings of Grb7 interacting with the 5′-untranslated region of capped targeted KOR (kappa opioid receptor) mRNA and the Hu antigen R of stress granules in an FAK-mediated phosphorylation manner (24, 25).Unlike its member proteins Grb10 and Grb14, the role of Grb7 in cell migration is unambiguous and well documented. This is supported by a series of studies. Firstly, Grb7 family members share a significantly conserved molecular architecture with the Caenorhabditis elegans Mig-10 protein, which is involved in neuronal cell migration during embryonic development (4, 5, 26), suggesting that Grb7 may play a role in cell migration. Moreover, Grb7 is often co-amplified with Her2/ErbB2 in certain human cancers and tumor cell lines (7, 27, 28), and its overexpression resulted in invasive and metastatic consequences of various cancers and tumor cells (23, 2933). On the contrary, knocking down Grb7 by RNA interference conferred to an inhibitory outcome of the breast cancer motility (34). Furthermore, interaction of Grb7 with autophosphorylated FAK at Tyr-397 could promote integrin-mediated cell migration in NIH 3T3 and CHO cells, whereas overexpression of its SH2 domain, an dominant negative mutant of Grb7, inhibited cell migration (19, 35). Recruitment and phosphorylation of Grb7 by EphB1 receptors enhanced cell migration in an ephrin-dependent manner (16). Recently, G7–18NATE, a selective Grb7-SH2 domain affinity cyclic peptide, was demonstrated to efficiently block cell migration of tumor cells (32, 36). In addition to cell migration, Grb7 has been shown to play a role in a variety of physiological and pathological events, for instance, kidney development (37), tumorigenesis (7, 14, 3841), angiogenic activity (20), proliferation (34, 42, 43), anti-apoptosis (44), gene expression regulation (24), Silver-Russell syndrome (45), rheumatoid arthritis (46), atopic dermatitis (47), and T-cell activation (17, 48). Nevertheless, it remains largely unknown regarding the downstream signaling events of Grb7-mediated various functions. In particular, given the role of Grb7 as an adaptor molecule and its SH2 domain mainly interacting with upstream regulators, it will be interesting to identify potential downstream effectors through interacting with the functional GM region or N-terminal proline-rich region.In this report, we identified two tyrosine phosphorylated sites of Grb7 by FAK and deciphered the signaling targets downstream through these phosphorylated tyrosine sites to regulate various cellular functions such as cell migration, proliferation, and survival. In addition, our study sheds light on tyrosine phosphorylation of Grb7 by FAK involved in tumorigenesis.  相似文献   

11.
Early onset generalized dystonia (DYT1) is an autosomal dominant neurological disorder caused by deletion of a single glutamate residue (torsinA ΔE) in the C-terminal region of the AAA+ (ATPases associated with a variety of cellular activities) protein torsinA. The pathogenic mechanism by which torsinA ΔE mutation leads to dystonia remains unknown. Here we report the identification and characterization of a 628-amino acid novel protein, printor, that interacts with torsinA. Printor co-distributes with torsinA in multiple brain regions and co-localizes with torsinA in the endoplasmic reticulum. Interestingly, printor selectively binds to the ATP-free form but not to the ATP-bound form of torsinA, supporting a role for printor as a cofactor rather than a substrate of torsinA. The interaction of printor with torsinA is completely abolished by the dystonia-associated torsinA ΔE mutation. Our findings suggest that printor is a new component of the DYT1 pathogenic pathway and provide a potential molecular target for therapeutic intervention in dystonia.Early onset generalized torsion dystonia (DYT1) is the most common and severe form of hereditary dystonia, a movement disorder characterized by involuntary movements and sustained muscle spasms (1). This autosomal dominant disease has childhood onset and its dystonic symptoms are thought to result from neuronal dysfunction rather than neurodegeneration (2, 3). Most DYT1 cases are caused by deletion of a single glutamate residue at positions 302 or 303 (torsinA ΔE) of the 332-amino acid protein torsinA (4). In addition, a different torsinA mutation that deletes amino acids Phe323–Tyr328 (torsinA Δ323–328) was identified in a single family with dystonia (5), although the pathogenic significance of this torsinA mutation is unclear because these patients contain a concomitant mutation in another dystonia-related protein, ϵ-sarcoglycan (6). Recently, genetic association studies have implicated polymorphisms in the torsinA gene as a genetic risk factor in the development of adult-onset idiopathic dystonia (7, 8).TorsinA contains an N-terminal endoplasmic reticulum (ER)3 signal sequence and a 20-amino acid hydrophobic region followed by a conserved AAA+ (ATPases associated with a variety of cellular activities) domain (9, 10). Because members of the AAA+ family are known to facilitate conformational changes in target proteins (11, 12), it has been proposed that torsinA may function as a molecular chaperone (13, 14). TorsinA is widely expressed in brain and multiple other tissues (15) and is primarily associated with the ER and nuclear envelope (NE) compartments in cells (1620). TorsinA is believed to mainly reside in the lumen of the ER and NE (1719) and has been shown to bind lamina-associated polypeptide 1 (LAP1) (21), lumenal domain-like LAP1 (LULL1) (21), and nesprins (22). In addition, recent evidence indicates that a significant pool of torsinA exhibits a topology in which the AAA+ domain faces the cytoplasm (20). In support of this topology, torsinA is found in the cytoplasm, neuronal processes, and synaptic terminals (2, 3, 15, 2326) and has been shown to bind cytosolic proteins snapin (27) and kinesin light chain 1 (20). TorsinA has been proposed to play a role in several cellular processes, including dopaminergic neurotransmission (2831), NE organization and dynamics (17, 22, 32), and protein trafficking (27, 33). However, the precise biological function of torsinA and its regulation remain unknown.To gain insights into torsinA function, we performed yeast two-hybrid screens to search for torsinA-interacting proteins in the brain. We report here the isolation and characterization of a novel protein named printor (protein interactor of torsinA) that interacts selectively with wild-type (WT) torsinA but not the dystonia-associated torsinA ΔE mutant. Our data suggest that printor may serve as a cofactor of torsinA and provide a new molecular target for understanding and treating dystonia.  相似文献   

12.
Paneth cells are a secretory epithelial lineage that release dense core granules rich in host defense peptides and proteins from the base of small intestinal crypts. Enteric α-defensins, termed cryptdins (Crps) in mice, are highly abundant in Paneth cell secretions and inherently resistant to proteolysis. Accordingly, we tested the hypothesis that enteric α-defensins of Paneth cell origin persist in a functional state in the mouse large bowel lumen. To test this idea, putative Crps purified from mouse distal colonic lumen were characterized biochemically and assayed in vitro for bactericidal peptide activities. The peptides comigrated with cryptdin control peptides in acid-urea-PAGE and SDS-PAGE, providing identification as putative Crps. Matrix-assisted laser desorption ionization time-of-flight mass spectrometry experiments showed that the molecular masses of the putative α-defensins matched those of the six most abundant known Crps, as well as N-terminally truncated forms of each, and that the peptides contain six Cys residues, consistent with identities as α-defensins. N-terminal sequencing definitively revealed peptides with N termini corresponding to full-length, (des-Leu)-truncated, and (des-Leu-Arg)-truncated N termini of Crps 1–4 and 6. Crps from mouse large bowel lumen were bactericidal in the low micromolar range. Thus, Paneth cell α-defensins secreted into the small intestinal lumen persist as intact and functional forms throughout the intestinal tract, suggesting that the peptides may mediate enteric innate immunity in the colonic lumen, far from their upstream point of secretion in small intestinal crypts.Antimicrobial peptides (AMPs)2 are released by epithelial cells onto mucosal surfaces as effectors of innate immunity (15). In mammals, most AMPs derive from two major families, the cathelicidins and defensins (6). The defensins comprise the α-, β-, and θ-defensin subfamilies, which are defined by the presence of six cysteine residues paired in characteristic tridisulfide arrays (7). α-Defensins are highly abundant in two primary cell lineages: phagocytic leukocytes, primarily neutrophils, of myeloid origin and Paneth cells, which are secretory epithelial cells located at the base of the crypts of Lieberkühn in the small intestine (810). Neutrophil α-defensins are stored in azurophilic granules and contribute to non-oxidative microbial cell killing in phagolysosomes (11, 12), except in mice whose neutrophils lack defensins (13). In the small bowel, α-defensins and other host defense proteins (1418) are released apically as components of Paneth cell secretory granules in response to cholinergic stimulation and after exposure to bacterial antigens (19). Therefore, the release of Paneth cell products into the crypt lumen is inferred to protect mitotically active crypt cells from colonization by potential pathogens and confer protection against enteric infection (7, 20, 21).Under normal, homeostatic conditions, Paneth cells are not found outside the small bowel, although they may appear ectopically in response to local inflammation throughout the gastrointestinal tract (22, 23). Paneth cell numbers increase progressively throughout the small intestine, occurring at highest numbers in the distal ileum (24). Mouse Paneth cells express numerous α-defensin isoforms, termed cryptdins (Crps) (25), that have broad spectrum antimicrobial activities (6, 26). Collectively, α-defensins constitute approximately seventy percent of the bactericidal peptide activity in mouse Paneth cell secretions (19), selectively killing bacteria by membrane-disruptive mechanisms (2730). The role of Paneth cell α-defensins in gastrointestinal mucosal immunity is evident from studies of mice transgenic for human enteric α-defensin-5, HD-5, which are immune to infection by orally administered Salmonella enterica sv. typhimurium (S. typhimurium) (31).The biosynthesis of mature, bactericidal α-defensins from their inactive precursors requires activation by lineage-specific proteolytic convertases. In mouse Paneth cells, inactive ∼8.4-kDa Crp precursors are processed intracellularly into microbicidal ∼4-kDa Crps by specific cleavage events mediated by matrix metalloproteinase-7 (MMP-7) (32, 33). MMP-7 null mice exhibit increased susceptibility to systemic S. typhimurium infection and decreased clearance of orally administered non-invasive Escherichia coli (19, 32). Although the α-defensin proregions are sensitive to proteolysis, the mature, disulfide-stabilized peptides resist digestion by their converting enzymes in vitro, whether the convertase is MMP-7 (32), trypsin (34), or neutrophil serine proteinases (35). Because α-defensins resist proteolysis in vitro, we hypothesized that Paneth cell α-defensins resist degradation and remain in a functional state in the large bowel, a complex, hostile environment containing varied proteases of both host and microbial origin.Here, we report on the isolation and characterization of a population of enteric α-defensins from the mouse colonic lumen. Full-length and N-terminally truncated Paneth cell α-defensins were identified and are abundant in the distal large bowel lumen.  相似文献   

13.
The molecular chaperone Hsp27 exists as a distribution of large oligomers that are disassembled by phosphorylation at Ser-15, -78, and -82. It is controversial whether the unphosphorylated Hsp27 or the widely used triple Ser-to-Asp phospho-mimic mutant is the more active molecular chaperone in vitro. This question was investigated here by correlating chaperone activity, as measured by the aggregation of reduced insulin or α-lactalbumin, with Hsp27 self-association as monitored by analytical ultracentrifugation. Furthermore, because the phospho-mimic is generally assumed to reproduce the phosphorylated molecule, the size and chaperone activity of phosphorylated Hsp27 were compared with that of the phospho-mimic. Hsp27 was triply phosphorylated by MAPKAP-2 kinase, and phosphorylation was tracked by urea-PAGE. An increasing degree of suppression of insulin or α-lactalbumin aggregation correlated with a decreasing Hsp27 self-association, which was the least for phosphorylated Hsp27 followed by the mimic followed by the unphosphorylated protein. It was also found that Hsp27 added to pre-aggregated insulin did not reverse aggregation but did inhibit these aggregates from assembling into even larger aggregates. This chaperone activity appears to be independent of Hsp27 phosphorylation. In conclusion, the most active chaperone of insulin and α-lactalbumin was the Hsp27 (elongated) dimer, the smallest Hsp27 subunit observed under physiological conditions. Next, the Hsp27 phospho-mimic is only a partial mimic of phosphorylated Hsp27, both in self-association and in chaperone function. Finally, the efficient inhibition of insulin aggregation by Hsp27 dimer led to the proposal of two models for this chaperone activity.Oligomeric heat shock protein 27 (Hsp27)2 is a ubiquitous mammalian protein with a variety of functions in health and disease (18). These functions include ATP-independent chaperone activity in response to environmental stress, e.g. heat shock and oxidative stress, control of apoptosis, and regulation of actin cytoskeleton dynamics. Hsp27 is a member of the α-crystallin small heat shock protein family of which αB-crystallin is the archetype. These proteins are characterized by an α-crystallin domain of 80–90 residues consisting of roughly eight β-strands that form an intermolecular β-sheet interaction interface within a dimer, the basic building subunit of the oligomer (2, 4, 911).Hsp27 is in equilibrium between high molecular weight oligomers and much lower molecular weight multimers. It has been reported that unphosphorylated Hsp27 includes predominantly a distribution of high molecular species ranging in size from 12-mer to 35-mer (1219). Phosphorylation of Hsp27 at serines 15, 78, and 82 by the p38-activated MAPKAP-2 kinase (2022) or the use of the triple Ser-to-Asp phospho-mimic results in a major shift in the equilibrium toward much smaller multimers (23) and in an alteration of its function (1, 3, 6, 7, 24, 25). The size distribution of the smaller species has been reported to be between monomer and tetramer (1216, 18, 19).Small heat shock proteins, including Hsp27, behave as ATP-independent molecular chaperones during cellular heat shock. They bind partially unfolded proteins and prevent their aggregation until the proteins can be refolded by larger ATP-dependent chaperones or are digested (7, 8, 26). This function includes the up-regulation and/or phosphorylation of Hsp27.It is not entirely clear what the role of Hsp27 size and phosphorylation state plays in its heat shock function because there are conflicting results in the literature. Some in vitro studies concluded that the unphosphorylated oligomeric Hsp27 (or the murine isoform Hsp25) protects proteins against aggregation better than does the phosphorylation mimic (13, 19, 27), whereas others found no difference (16, 28, 29), and still other studies found that the mimic protects better than does the unphosphorylated wild type (27, 30, 31). In-cell studies found that phosphorylation of Hsp27 was essential for thermo-protection of actin filaments (32), and the Hsp27 phosphorylation mimic decreased inclusion body formation better than did unphosphorylated Hsp27 (33). This study was undertaken to investigate the molecular chaperone function of Hsp27 by correlating chaperone activity with Hsp27 size and by comparing fully phosphorylated Hsp27 with its phospho-mimic.  相似文献   

14.
15.
16.
17.
Prion propagation involves a conformational transition of the cellular form of prion protein (PrPC) to a disease-specific isomer (PrPSc), shifting from a predominantly α-helical conformation to one dominated by β-sheet structure. This conformational transition is of critical importance in understanding the molecular basis for prion disease. Here, we elucidate the conformational properties of a disulfide-reduced fragment of human PrP spanning residues 91–231 under acidic conditions, using a combination of heteronuclear NMR, analytical ultracentrifugation, and circular dichroism. We find that this form of the protein, which similarly to PrPSc, is a potent inhibitor of the 26 S proteasome, assembles into soluble oligomers that have significant β-sheet content. The monomeric precursor to these oligomers exhibits many of the characteristics of a molten globule intermediate with some helical character in regions that form helices I and III in the PrPC conformation, whereas helix II exhibits little evidence for adopting a helical conformation, suggesting that this region is a likely source of interaction within the initial phases of the transformation to a β-rich conformation. This precursor state is almost as compact as the folded PrPC structure and, as it assembles, only residues 126–227 are immobilized within the oligomeric structure, leaving the remainder in a mobile, random-coil state.Prion diseases, such as Creutzfeldt-Jacob and Gerstmann-Sträussler-Scheinker in humans, scrapie in sheep, and bovine spongiform encephalopathy in cattle, are fatal neurological disorders associated with the deposition of an abnormally folded form of a host-encoded glycoprotein, prion (PrP)2 (1). These diseases may be inherited, arise sporadically, or be acquired through the transmission of an infectious agent (2, 3). The disease-associated form of the protein, termed the scrapie form or PrPSc, differs from the normal cellular form (PrPC) through a conformational change, resulting in a significant increase in the β-sheet content and protease resistance of the protein (3, 4). PrPC, in contrast, consists of a predominantly α-helical structured domain and an unstructured N-terminal domain, which is capable of binding a number of divalent metals (512). A single disulfide bond links two of the main α-helices and forms an integral part of the core of the structured domain (13, 14).According to the protein-only hypothesis (15), the infectious agent is composed of a conformational isomer of PrP (16) that is able to convert other isoforms to the infectious isomer in an autocatalytic manner. Despite numerous studies, little is known about the mechanism of conversion of PrPC to PrPSc. The most coherent and general model proposed thus far is that PrPC fluctuates between the dominant native state and minor conformations, one or a set of which can self-associate in an ordered manner to produce a stable supramolecular structure composed of misfolded PrP monomers (3, 17). This stable, oligomeric species can then bind to, and stabilize, rare non-native monomer conformations that are structurally complementary. In this manner, new monomeric chains are recruited and the system can propagate.In view of the above model, considerable effort has been devoted to generating and characterizing alternative, possibly PrPSc-like, conformations in the hope of identifying common properties or features that facilitate the formation of amyloid oligomers. This has been accomplished either through PrPSc-dependent conversion reactions (1820) or through conversion of PrPC in the absence of a PrPSc template (2125). The latter approach, using mainly disulfide-oxidized recombinant PrP, has generated a wide range of novel conformations formed under non-physiological conditions where the native state is relatively destabilized. These conformations have ranged from near-native (14, 26, 27), to those that display significant β-sheet content (21, 23, 2833). The majority of these latter species have shown a high propensity for aggregation, although not all are on-pathway to the formation of amyloid. Many of these non-native states also display some of the characteristics of PrPSc, such as increased β-sheet content, protease resistance, and a propensity for oligomerization (28, 29, 31) and some have been claimed to be associated with the disease process (34).One such PrP folding intermediate, termed β-PrP, differs from the majority of studied PrP intermediate states in that it is formed by refolding the PrP molecule from the native α-helical conformation (here termed α-PrP), at acidic pH in a reduced state, with the disulfide bond broken (22, 35). Although no covalent differences between the PrPC and PrPSc have been consistently identified to date, the role of the disulfide bond in prion propagation remains disputed (25, 3639). β-PrP is rich in β-sheet structure (22, 35), and displays many of the characteristics of a PrPSc-like precursor molecule, such as partial resistance to proteinase K digestion, and the ability to form amyloid fibrils in the presence of physiological concentrations of salts (40).The β-PrP species previously characterized, spanning residues 91–231 of PrP, was soluble at low ionic strength buffers and monomeric, according to elution volume on gel filtration (22). NMR analysis showed that it displayed radically different spectra to those of α-PrP, with considerably fewer observable peaks and markedly reduced chemical shift dispersion. Data from circular dichroism experiments showed that fixed side chain (tertiary) interactions were lost, in contrast to the well defined β-sheet secondary structure, and thus in conjunction with the NMR data, indicated that β-PrP possessed a number of characteristics associated with a “molten globule” folding intermediate (22). Such states have been proposed to be important in amyloid and fibril formation (41). Indeed, antibodies raised against β-PrP (e.g. ICSM33) are capable of recognizing native PrPSc (but not PrPC) (4244). Subsequently, a related study examining the role of the disulfide bond in PrP folding confirmed that a monomeric molten globule-like form of PrP was formed on refolding the disulfide-reduced protein at acidic pH, but reported that, under their conditions, the circular dichroism response interpreted as β-sheet structure was associated with protein oligomerization (45). Indeed, atomic force microscopy on oligomeric full-length β-PrP (residues 23–231) shows small, round particles, showing that it is capable of formation of oligomers without forming fibrils (35). Notably, however, salt-induced oligomeric β-PrP has been shown to be a potent inhibitor of the 26 S proteasome, in a similar manner to PrPSc (46). Impairment of the ubiquitin-proteasome system in vivo has been linked to prion neuropathology in prion-infected mice (46).Although the global properties of several PrP intermediate states have been determined (3032, 35), no information on their conformational properties on a sequence-specific basis has been obtained. Their conformational properties are considered important, as the elucidation of the chain conformation may provide information on the way in which these chains pack in the assembly process, and also potentially provide clues on the mechanism of amyloid assembly and the phenomenon of prion strains. As the conformational fluctuations and heterogeneity of molten globule states give rise to broad NMR spectra that preclude direct observation of their conformational properties by NMR (4750), here we use denaturant titration experiments to determine the conformational properties of β-PrP, through the population of the unfolded state that is visible by NMR. In addition, we use circular dichroism and analytical ultracentrifugation to examine the global structural properties, and the distribution of multimeric species that are formed from β-PrP.  相似文献   

18.
19.
20.
Lysophosphatidic acid (LPA), a bioactive phospholipid, induces a wide range of cellular effects, including gene expression, cytoskeletal rearrangement, and cell survival. We have previously shown that LPA stimulates secretion of pro- and anti-inflammatory cytokines in bronchial epithelial cells. This study provides evidence that LPA enhances pulmonary epithelial barrier integrity through protein kinase C (PKC) δ- and ζ-mediated E-cadherin accumulation at cell-cell junctions. Treatment of human bronchial epithelial cells (HBEpCs) with LPA increased transepithelial electrical resistance (TER) by ∼2.0-fold and enhanced accumulation of E-cadherin to the cell-cell junctions through Gαi-coupled LPA receptors. Knockdown of E-cadherin with E-cadherin small interfering RNA or pretreatment with EGTA (0.1 mm) prior to LPA (1 μm) treatment attenuated LPA-induced increases in TER in HBEpCs. Furthermore, LPA induced tyrosine phosphorylation of focal adhesion kinase (FAK) and overexpression of the FAK inhibitor, and FAK-related non-kinase-attenuated LPA induced increases in TER and E-cadherin accumulation at cell-cell junctions. Overexpression of dominant negative protein kinase δ and ζ attenuated LPA-induced phosphorylation of FAK, accumulation of E-cadherin at cell-cell junctions, and an increase in TER. Additionally, lipopolysaccharide decreased TER and induced E-cadherin relocalization from cell-cell junctions to cytoplasm in a dose-dependent fashion, which was restored by LPA post-treatment in HBEpCs. Intratracheal post-treatment with LPA (5 μm) reduced LPS-induced neutrophil influx, protein leak, and E-cadherin shedding in bronchoalveolar lavage fluids in a murine model of acute lung injury. These data suggest a protective role of LPA in airway inflammation and remodeling.The airway epithelium is the site of first contact for inhaled environmental stimuli, functions as a physical barrier to environmental insult, and is an essential part of innate immunity. Epithelial barrier disruption is caused by inhaled allergens, dust, and irritants, resulting in inflammation, bronchoconstriction, and edema as seen in asthma and other respiratory diseases (14). Furthermore, increased epithelial permeability also results in para-cellular leakage of large proteins, such as albumin, immunoglobulin G, and polymeric immunoglobulin A, into the airway lumen (5, 6). The epithelial cell-cell junctional complex is composed of tight junctions, adherens junctions, and desmosomes. These adherens junctions play a pivotal role in regulating the activity of the entire junctional complex because the formation of adherens junctions subsequently leads to the formation of other cell-cell junctions (79). The major adhesion molecules in the adherens junctions are the cadherins. E-cadherin is a member of the cadherin family that mediates calcium-dependent cell-cell adhesion. The N-terminal ectodomain of E-cadherin contains homophilic interaction specificity, and the cytoplasmic domain binds to catenins, which interact with actin (1013). Plasma membrane localization of E-cadherin is critical for the maintenance of epithelial cell-cell junctions and airway epithelium integrity (7, 10, 14). A decrease of adhesive properties of E-cadherin is related to the loss of differentiation and the subsequent acquisition of a higher motility and invasiveness of epithelial cells (10, 14, 15). Dislocation or shedding of E-cadherin in the airway epithelium induces epithelial shedding and increases airway permeability in lung airway diseases (10, 14, 16). In an ovalbumin-challenged guinea pig model of asthma, it has been demonstrated that E-cadherin is dislocated from the lateral margins of epithelial cells (10). Histamine increases airway para-cellular permeability and results in an increased susceptibility of airway epithelial cells to adenovirus infection by interrupting E-cadherin adhesion (14). Serine phosphorylation of E-cadherin by casein kinase II, GSK-3β, and PKD1/PKC2 μ enhanced E-cadherin-mediated cell-cell adhesion in NIH3T3 fibroblasts and LNCaP prostate cancer cells (11, 17). However, the regulation and mechanism by which E-cadherin is localized within the pulmonary epithelium is not fully known, particularly during airway remodeling.LPA, a naturally occurring bioactive lipid, is present in body fluids, such as plasma, saliva, follicular fluid, malignant effusions, and bronchoalveolar lavage (BAL) fluids (1820). Six distinct high affinity cell-surface LPA receptors, LPA-R1–6, have been cloned and described in mammals (2126). Extracellular activities of LPA include cell proliferation, motility, and cell survival (2730). LPA exhibits a wide range of effects on differing cell types, including pulmonary epithelial, smooth muscle, fibroblasts, and T cells (3135). LPA augments migration and cytokine synthesis in lymphocytes and induces chemotaxis of Jurkat T cells through Matrigel membranes (34). LPA induces airway smooth muscle cell contractility, proliferation, and airway repair and remodeling (35, 36). LPA also potently stimulates IL-8 (31, 3739), IL-13 receptor α2 (IL-13Rα2) (40), and COX-2 gene expression and prostaglandin E2 release (41) in HBEpCs. Prostaglandin E2 and IL-13Rα2 have anti-inflammatory properties in pulmonary inflammation (42, 43). These results suggest that LPA may play a protective role in lung disease by stimulating an innate immune response while simultaneously attenuating the adaptive immune response. Furthermore, intravenous injection with LPA attenuated bacterial endotoxin-induced plasma tumor necrosis factor-α production and myeloperoxidase activity in the lungs of mice (44), suggesting an anti-inflammatory role for LPA in a murine model of sepsis.We have reported that LPA induces E-cadherin/c-Met accumulation in cell-cell contacts and increases TER in HBEpCs (45). Here, for the first time, we report that LPA-induced increases in TER are dependent on PKCδ, PKCζ, and FAK-mediated E-cadherin accumulation at cell-cell junctions. Furthermore, we demonstrate that post-treatment of LPA rescues LPS-induced airway epithelial disruption in vitro and reduces E-cadherin shedding in a murine model of ALI. This study identifies the molecular mechanisms linking the LPA and LPA receptors to maintaining normal pulmonary epithelium barrier function, which is critical in developing novel therapies directed at ameliorating pulmonary diseases.  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