首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The double-headed arrowhead protease inhibitors API-A and -B from the tubers of Sagittaria sagittifolia (Linn) feature two distinct reactive sites, unlike other members of their family. Although the two inhibitors have been extensively characterized, the identities of the two P1 residues in both API-A and -B remain controversial. The crystal structure of a ternary complex at 2.48 Å resolution revealed that the two trypsins bind on opposite sides of API-A and are 34 Å apart. The overall fold of API-A belongs to the β-trefoil fold and resembles that of the soybean Kunitz-type trypsin inhibitors. The two P1 residues were unambiguously assigned as Leu87 and Lys145, and their identities were further confirmed by site-directed mutagenesis. Reactive site 1, composed of residues P5 Met83 to P5′ Ala92, adopts a novel conformation with the Leu87 completely embedded in the S1 pocket even though it is an unfavorable P1 residue for trypsin. Reactive site 2, consisting of residues P5 Cys141 to P5′ Glu150, binds trypsin in the classic mode by employing a two-disulfide-bonded loop. Analysis of the two binding interfaces sheds light on atomic details of the inhibitor specificity and also promises potential improvements in enzyme activity by engineering of the reactive sites.Protease inhibitors (PIs)4 are ubiquitously distributed in all organisms, including plants, animals, and microorganisms (1). They play vital roles in regulating their corresponding proteases, which are involved in many biological processes such as protein digestion, cell signal transmission, inflammation, apoptosis, blood coagulation, and embryogenesis (2). The clinical applications of PIs are widespread, and there is great interest in developing more potent therapeutic PIs for treating human diseases related to cancer (3), pancreatitis (4), thrombosis (5), and AIDS (6). To this end, the soybean Kunitz-type serine proteases inhibitors have been extensively studied (1, 711). The inhibitors of this family generally contain 170–200 residues and have two disulfide bonds. Most members have only one reactive site located in the region of residues 60–70 (7, 10, 1214). However, a few members possess two reactive sites that simultaneously bind two protease molecules and are thus termed double-headed inhibitors (1518). All of these inhibitors are classified into family I3 of peptidase inhibitors (19). Most members are further grouped into subfamily I3A. However, the double-headed arrowhead PIs API-A and -B are grouped in subfamily I3B because of their very low sequence similarity to other members (19). In contrast to other double-headed PIs such as the Bowman-Birk and ovomucoid inhibitors, which have two identical reactive sites that have evolved by domain shuffling and gene duplication (1, 2025), both API-A and -B have two distinct reactive sites.API-A and -B were first purified from the tubers of Sagittaria sagittifolia (Linn) in 1979 (26). Both consist of 179 residues with three disulfide bonds and can inhibit a variety of serine proteases, including trypsin, chymotrypsin, and porcine tissue kallikrein (17, 2628). Although the sequence identity of API-A and -B is as high as 91%, their inhibitory specificities differ. The former can bind one molecule of trypsin and one molecule of chymotrypsin, whereas the latter can simultaneously bind two molecules of trypsin (26). The two P1 residues of the reactive sites of API-A and -B were first predicted to be Lys44 and Arg76 based on their surrounding sequences, which are similar to those of the reactive sites of bovine pancreas trypsin inhibitor and soybean Kunitz trypsin inhibitor (29). However, their identities were later revised to Arg76 and Leu87 (for API-A) or Lys87 (for API-B) based on results from sited-directed mutagenesis studies (30).To clarify these controversies, we solved the crystal structure of API-A in complex with two molecules of bovine trypsin. To the best of our knowledge, this is the first report on the three-dimensional structure of the double-headed Kunitz-type trypsin inhibitor in complex with two molecules of protease. On the basis of this structure, the two P1 residues have now been identified as Leu87 and Lys145 for reactive site 1 (RS1) and 2 (RS2), respectively. The results were further confirmed by site-directed mutagenesis. It was earlier shown that the first P1 residue Leu87 interacts preferentially with chymotrypsin (30). In our structure, Leu87 is snugly embedded in the S1 pocket of trypsin, as a consequence of the broad interface contributed by the surrounding residues. Comprehensive analyses of the two reactive site interfaces have provided functional insights into the novel inhibitory patterns of this unique double-headed protease inhibitor.  相似文献   

2.
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   

3.
Human FGF1 (fibroblast growth factor 1) is a powerful signaling molecule with a short half-life in vivo and a denaturation temperature close to physiological. Binding to heparin increases the stability of FGF1 and is believed to be important in the formation of FGF1·fibroblast growth factor receptor (FGFR) active complex. In order to reveal the function of heparin in FGF1·FGFR complex formation and signaling, we constructed several FGF1 variants with reduced affinity for heparin and with diverse stability. We determined their biophysical properties and biological activities as well as their ability to translocate across cellular membranes. Our study showed that increased thermodynamic stability of FGF1 nicely compensates for decreased binding of heparin in FGFR activation, induction of DNA synthesis, and cell proliferation. By stepwise introduction of stabilizing mutations into the K118E (K132E) FGF1 variant that shows reduced affinity for heparin and is inactive in stimulation of DNA synthesis, we were able to restore the full mitogenic activity of this mutant. Our results indicate that the main role of heparin in FGF-induced signaling is to protect this naturally unstable protein against heat and/or proteolytic degradation and that heparin is not essential for a direct FGF1-FGFR interaction and receptor activation.FGF1 (fibroblast growth factor 1) belongs to a family of polypeptide growth factors comprising in humans 22 structurally related proteins (1, 2). The signaling induced by the growth factor leads to a wide range of cellular responses during development as well as in adult life, such as growth regulation, differentiation, survival, stress response, migration, and proliferation of different cell types (3). The biological activity of FGF1 is exerted through binding to four high affinity cell surface receptors (FGFR1–4), resulting in receptor dimerization and transphosphorylation in its tyrosine kinase domain (4, 5). The activated FGFR3 induces cellular response by initiating several signaling cascades, including mitogen-activated protein kinase (MAPK), phosphoinositide 3-kinase/Akt, and phospholipase C-γ (PLC-γ) pathways (6).In addition to FGFRs, FGF1 binds to heparan sulfates (HS) associated with proteoglycans at the cell surface and in the extracellular matrix (7). Among the physiological sugars, the highest affinity for FGF1 is shown by heparin, a widely used linear, highly sulfated polysaccharide composed of 2-O-sulfated iduronic acid and 6-O-sulfated, N-sulfated glucosamine units (8).Despite many years of research, there is still controversy regarding the molecular role of heparin/HS in FGF1- and FGF2-induced signaling. Thus, the question of whether or not the linkage of two molecules of the growth factor by heparin/HS is an absolute prerequisite for induction of FGFR dimerization is still open. Numerous studies have concluded that the presence of heparin/HS is obligatory for FGF signaling. It is widely believed that heparin/HS is directly involved in receptor dimerization and is critical for mitogenic response stimulated by the growth factor (4, 6, 810).On the other hand, several authors working on FGF1 and FGF2 have suggested that there is no mandatory requirement for heparin for the assembly and activation of the FGF·FGFR complex. They imply that heparin only plays a role in association of two molecules of the growth factor and therefore facilitates their binding to FGFR (11). It has been reported that FGF1 and FGF2 can interact with the FGFR and trigger phosphorylation of p42/44 MAPK and activation of other signaling pathways even in the absence of HS (1216).The accepted role of heparin/HS in FGF1 signaling is to prevent the degradation of the growth factor (17). The interaction with heparin or HS protects FGF1 against heat, acidic pH, and proteases (18, 19). HS also seems to regulate the activity of different FGFs by creating their local reservoir and generating a concentration gradient of the growth factor (6, 17).The binding of FGF1 to heparin/HS is mediated by specific residues forming a positively charged patch on the protein surface (20, 21). The major contribution is made by Lys118 (Lys132 in the full-length numbering system), which was identified by Harper and Lobb (22), and Lys112 and Arg122 (23, 24). Additional residues of FGF1 involved in the interaction with heparin are the positively charged Lys113, Arg119, and Lys128 and the polar Asn18, Asn114, and Gln127 (20, 21). Site-directed mutagenesis and other studies have revealed the importance of Lys118 not only in heparin binding but also for the biological function of FGF1 (22, 25, 26). It was shown that the K118E (K132E) mutant is inactive in stimulation of DNA synthesis, although its affinity for FGFR and the ability to activate signaling cascades is not reduced (27, 28). Despite extensive research, the reason for the lack of mitogenic potential of K118E FGF1 is still not clear.In this paper, we verified the function of heparin in FGF1·FGFR complex formation and signaling by constructing several FGF1 mutants with reduced affinity for heparin. To recover the stability of these variants, which could no longer be stabilized by heparin, we supplemented them stepwise with stabilizing mutations (29). We analyzed thoroughly their biological activity and their ability to translocate across cellular membranes (3034). Interestingly, the full mitogenic activity of the K118E FGF1 variant was restored by the introduced stabilizing mutations.Our results indicate that the main role of heparin in FGF-induced signaling is to protect this naturally unstable protein against heat denaturation and proteolytic degradation and that the increased stability of the growth factor can compensate for reduced heparin binding.  相似文献   

4.
In replicating yeast, lysine 63-linked polyubiquitin (polyUb) chains are extended from the ubiquitin moiety of monoubiquitinated proliferating cell nuclear antigen (monoUb-PCNA) by the E2-E3 complex of (Ubc13-Mms2)-Rad5. This promotes error-free bypass of DNA damage lesions. The unusual ability of Ubc13-Mms2 to synthesize unanchored Lys63-linked polyUb chains in vitro allowed us to resolve the individual roles that it and Rad5 play in the catalysis and specificity of PCNA polyubiquitination. We found that Rad5 stimulates the synthesis of free polyUb chains by Ubc13-Mms2 in part by enhancing the reactivity of the Ubc13∼Ub thiolester bond. Polyubiquitination of monoUb-PCNA was further enhanced by interactions between the N-terminal domain of Rad5 and PCNA. Thus, Rad5 acts both to align monoUb-PCNA with Ub-charged Ubc13 and to stimulate Ub transfer onto Lys63 of a Ub acceptor. We also found that Rad5 interacts with PCNA independently of the number of monoubiquitinated subunits in the trimer and that it binds to both unmodified and monoUb-PCNA with similar affinities. These findings indicate that Rad5-mediated recognition of monoUb-PCNA in vivo is likely to depend upon interactions with additional factors at stalled replication forks.DNA is susceptible to chemical alteration by many endogenous and exogenous agents. To counter this threat and maintain genome integrity, eukaryotic cells employ three main strategies: DNA repair pathways that directly reverse DNA damage, cell cycle checkpoints that allow time to repair the damage prior to replication, and DNA damage tolerance (DDT),2 which is a method of bypassing DNA damage lesions during the DNA replication phase of the cell cycle.Proliferating cell nuclear antigen (PCNA) is a key regulatory protein in DNA replication and repair (1). At the replication fork, DNA is encircled by PCNA, a homotrimeric protein that promotes processive movement of the replicative DNA polymerase. Upon DNA damage and subsequent stalling of the replicative polymerase, Ub modifications of PCNA signal DDT, which allows a cell to bypass the lesion and proceed past this potential block in replication (24).In the DDT pathway, as in other Ub-dependent pathways, Ub is conjugated to a substrate by the actions of three enzymes, an E1 activating enzyme, an E2 conjugating enzyme, and an E3 ligase (5). The E1 enzyme initiates the pathway in a two-step reaction that utilizes ATP hydrolysis to activate the C terminus of Ub, culminating in the formation of an E1∼Ub thiolester. Subsequent transthiolation to the active site cysteine of the E2 generates an E2∼Ub thiolester. An E3 ligase then brings a substrate into close proximity to the E2∼Ub intermediate, thereby catalyzing the formation of an isopeptide bond between the amino group of a substrate lysine and the C-terminal glycine of Ub. Polyubiquitination occurs when this substrate is another Ub, either free or as part of a Ub-protein conjugate.The DDT pathway is characterized by distinct ubiquitination events on PCNA that occur in two stages (3, 4, 6). The first of these is monoubiquitination of lysine 164 on one or more of the PCNA subunits by the E2-E3 complex of Rad6-Rad18 in Saccharomyces cerevisiae (3, 4, 7). monoUb-PCNA can serve either as a signal for error-prone bypass of the DNA lesion by recruiting translesion polymerases or as a substrate for subsequent polyubiquitination by the E2 heterodimer Ubc13-Mms2 and the E3 ligase Rad5 (3, 4, 8, 9). The polyUb chain extended from the initial Ub moiety on monoUb-PCNA is linked specifically through Ub Lys63 residues. This Lys63-linked chain is thought to enable a template switch mechanism that allows for error-free bypass of the DNA lesion, in part by utilizing the single-strand DNA-dependent helicase activity of Rad5 (3, 4, 10, 11). Both PCNA ubiquitination events promote bypass of the DNA lesion rather than direct removal or repair of the lesion.We have been interested in the mechanism by which the yeast (Ubc13-Mms2)-Rad5 complex catalyzes the formation of Lys63-linked polyUb on PCNA. Previous studies have shown that heterodimerization of the Ubc13-Mms2 E2 is essential for Lys63-specific Ub-Ub conjugation in vitro and in vivo (1215). Ubc13 is a canonical E2 enzyme with an active site cysteine that receives activated Ub by transthiolation from the E1∼Ub complex (12, 13). This Ub is referred to as the “donor Ub.” Mms2 is a Ub E2 variant protein that lacks the active site cysteine (12, 15); rather, Mms2 binds to a second Ub, the “acceptor Ub,” and positions it to facilitate nucleophilic attack on the Ubc13∼Ub thiolester bond by the ϵ-amine of Lys63 (15, 16). The positioning of the acceptor Ub by Mms2 controls the specificity of polyUb assembly such that only Lys63-linked chains can be formed (16).Ubc13-Mms2 can synthesize Lys63-linked chains in vitro in the absence of a PCNA substrate or an E3 ligase (12, 13). However, unlike the synthesis of free Lys63-linked polyUb chains by Ubc13-Mms2, little is known about the polyubiquitination of PCNA or the role of the Rad5 E3 ligase in these reactions. Rad5 can bind PCNA and Rad18, and it contains a catalytic RING domain that characterizes the largest class of E3 ligases (1721). There is evidence that RING E3s like Rad5 may play a more active role in ubiquitination than simply bringing the substrate into close proximity with the E2∼Ub. Several RING E3s have been shown to stimulate the synthesis of unanchored polyUb chains or autoubiquitination of their cognate E2s in the absence of substrates (2224). This stimulation may be related to the ability of RING E3s to enhance reactivity of the E2∼Ub thiolester bond through allosteric effects (25, 26).Using purified recombinant forms of Ubc13, Mms2, and Rad5, we have explored the assembly of free Lys63-linked polyUb chains as well as the extension of a polyUb chain on a synthetic analog of monoUb-PCNA. We show that Rad5 facilitates ubiquitination in part by increasing the reactivity of the Ubc13∼Ub thiolester bond. With monoUb-PCNA substrates, Rad5 also stimulated polyubiquitination through direct interactions with PCNA and recruitment of Ub-charged Ubc13-Mms2. Surprisingly, Rad5 recognition of monoUb-PCNA appeared to depend on interactions only with the PCNA moiety of the conjugate, which suggests that substrate selectivity in vivo is likely to depend on additional factors.  相似文献   

5.
The flesh-eating bacterium group A Streptococcus (GAS) binds and activates human plasminogen, promoting invasive disease. Streptococcal surface enolase (SEN), a glycolytic pathway enzyme, is an identified plasminogen receptor of GAS. Here we used mass spectrometry (MS) to confirm that GAS SEN is octameric, thereby validating in silico modeling based on the crystal structure of Streptococcus pneumoniae α-enolase. Site-directed mutagenesis of surface-located lysine residues (SENK252 + 255A, SENK304A, SENK334A, SENK344E, SENK435L, and SENΔ434–435) was used to examine their roles in maintaining structural integrity, enzymatic function, and plasminogen binding. Structural integrity of the GAS SEN octamer was retained for all mutants except SENK344E, as determined by circular dichroism spectroscopy and MS. However, ion mobility MS revealed distinct differences in the stability of several mutant octamers in comparison with wild type. Enzymatic analysis indicated that SENK344E had lost α-enolase activity, which was also reduced in SENK334A and SENΔ434–435. Surface plasmon resonance demonstrated that the capacity to bind human plasminogen was abolished in SENK252 + 255A, SENK435L, and SENΔ434–435. The lysine residues at positions 252, 255, 434, and 435 therefore play a concerted role in plasminogen acquisition. This study demonstrates the ability of combining in silico structural modeling with ion mobility-MS validation for undertaking functional studies on complex protein structures.Streptococcus pyogenes (group A Streptococcus, GAS)8 is a common bacterial pathogen, causing over 700 million human disease episodes each year (1). These range from serious life-threatening invasive diseases including necrotizing fasciitis and streptococcal toxic shock-like syndrome to non-invasive infections like pharyngitis and pyoderma. Invasive disease, in combination with postinfection immune sequelae including rheumatic heart disease and acute poststreptococcal glomerulonephritis, account for over half a million deaths each year (1). Although a resurgence of GAS invasive infections has occurred in western countries since the mid-1980s, disease burden is much greater in developing countries and indigenous populations of developed nations, where GAS infections are endemic (24).GAS is able to bind human plasminogen and activate the captured zymogen to the serine protease plasmin (517). The capacity of GAS to do this plays a critical role in virulence and invasive disease initiation (3, 1719). The plasminogen activation system in humans is an important and highly regulated process that is responsible for breakdown of extracellular matrix components, dissolution of blood clots, and cell migration (20, 21). Plasminogen is a 92-kDa zymogen that circulates in human plasma at a concentration of 2 μm (22). It consists of a binding region of five homologous triple loop kringle domains and an N-terminal serine protease domain that flank the Arg561–Val562 site (23), where it is cleaved by tissue plasminogen activator and urokinase plasminogen activator to yield the active protease plasmin (20, 23). GAS also has the ability to activate human plasminogen by secreting the virulence determinant streptokinase. Streptokinase forms stable complexes with plasminogen or plasmin, both of which exhibit plasmin activity (20, 24). Activation of plasminogen by the plasmin(ogen)-streptokinase complex circumvents regulation by the host plasminogen activation inhibitors, α2-antiplasmin and α2-macroglobulin (11, 20). GAS can bind the plasmin(ogen)-streptokinase complex and/or plasmin(ogen) directly via plasmin(ogen) receptors at the bacterial cell surface (6). These receptors include the plasminogen-binding group A streptococcal M-like protein (PAM) (25), the PAM-related protein (19), glyceraldehyde-3-phosphate dehydrogenase (GAPDH; also known as streptococcal plasmin receptor, Plr, or streptococcal surface dehydrogenase) (9, 26), and streptococcal surface enolase (SEN or α-enolase) (27). Interactions with these GAS receptors occurs via lysine-binding sites within the kringle domains of plasminogen (6).In addition to its ability to bind human plasminogen, SEN is primarily the glycolytic enzyme that converts 2-phosphoglycerate to phosphoenolpyruvate (2729). SEN is abundantly expressed in the cytosol of most bacterial species but has also been identified as a surface-located protein in GAS and other bacteria including pneumococci, despite lacking classical cell surface protein motifs such as a signal sequence, membrane-spanning domain, or cell-wall anchor motif (27, 28, 30, 31). The interaction between SEN and plasminogen is reported to be facilitated by the two C-terminal lysine residues at positions 434 and 435 (27, 32). In contrast, an internal binding motif containing lysines at positions 252 and 255 in the closely related α-enolase of Streptococcus pneumoniae has been shown to play a pivotal role in the acquisition of plasminogen in this bacterial species (33). The octameric pneumococcal α-enolase structure consists of a tetramer of dimers. Hence, potential binding sites could be buried in the interface between subunits. In fact, the crystal structure of S. pneumoniae α-enolase revealed that the two C-terminal lysine residues are significantly less exposed than the internal plasminogen-binding motif (34).In this study, we constructed an in silico model of GAS SEN, based on the pneumococcal octameric α-enolase crystal structure, and validated this model using ion mobility (IM) mass spectrometry (MS). Site-directed mutagenesis followed by structural and functional analyses revealed that Lys344 plays a crucial role in structural integrity and enzymatic function. Furthermore, we demonstrate that the plasminogen-binding motif residues Lys252 and Lys255 and the C-terminal Lys434 and Lys435 residues are located adjacently in the GAS SEN structure and play a concerted role in the binding of human plasminogen.  相似文献   

6.
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   

7.
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   

8.
SLC26A7 (human)/Slc26a7 (mouse) is a recently identified chloride-base exchanger and/or chloride transporter that is expressed on the basolateral membrane of acid-secreting cells in the renal outer medullary collecting duct (OMCD) and in gastric parietal cells. Here, we show that mice with genetic deletion of Slc26a7 expression develop distal renal tubular acidosis, as manifested by metabolic acidosis and alkaline urine pH. In the kidney, basolateral Cl/HCO3 exchange activity in acid-secreting intercalated cells in the OMCD was significantly decreased in hypertonic medium (a normal milieu for the medulla) but was reduced only mildly in isotonic medium. Changing from a hypertonic to isotonic medium (relative hypotonicity) decreased the membrane abundance of Slc26a7 in kidney cells in vivo and in vitro. In the stomach, stimulated acid secretion was significantly impaired in isolated gastric mucosa and in the intact organ. We propose that SLC26A7 dysfunction should be investigated as a potential cause of unexplained distal renal tubular acidosis or decreased gastric acid secretion in humans.The collecting duct segment of the distal kidney nephron plays a major role in systemic acid base homeostasis by acid secretion and bicarbonate absorption. The acid secretion occurs via H+-ATPase and H-K-ATPase into the lumen and bicarbonate is absorbed via basolateral Cl/HCO3 exchangers (14). The tubules, which are located within the outer medullary region of the kidney collecting duct (OMCD),2 have the highest rate of acid secretion among the distal tubule segments and are therefore essential to the maintenance of acid base balance (2).The gastric parietal cell is the site of generation of acid and bicarbonate through the action of cytosolic carbonic anhydrase II (5, 6). The intracellular acid is secreted into the lumen via gastric H-K-ATPase, which works in conjunction with a chloride channel and a K+ recycling pathway (710). The intracellular bicarbonate is transported to the blood via basolateral Cl/HCO3 exchangers (1114).SLC26 (human)/Slc26 (mouse) isoforms are members of a conserved family of anion transporters that display tissue-specific patterns of expression in epithelial cells (1524). Several SLC26 members can function as chloride/bicarbonate exchangers. These include SLC26A3 (DRA), SLC26A4 (pendrin), SLC26A6 (PAT1 or CFEX), SLC26A7, and SLC26A9 (2531). SLC26A7 and SLC26A9 can also function as chloride channels (3234).SLC26A7/Slc26a7 is predominantly expressed in the kidney and stomach (28, 29). In the kidney, Slc26a7 co-localizes with AE1, a well-known Cl/HCO3 exchanger, on the basolateral membrane of (acid-secreting) A-intercalated cells in OMCD cells (29, 35, 36) (supplemental Fig. 1). In the stomach, Slc26a7 co-localizes with AE2, a major Cl/HCO3 exchanger, on the basolateral membrane of acid secreting parietal cells (28). To address the physiological function of Slc26a7 in the intact mouse, we have generated Slc26a7 ko mice. We report here that Slc26a7 ko mice exhibit distal renal tubular acidosis and impaired gastric acidification in the absence of morphological abnormalities in kidney or stomach.  相似文献   

9.
10.
11.
A decoding algorithm is tested that mechanistically models the progressive alignments that arise as the mRNA moves past the rRNA tail during translation elongation. Each of these alignments provides an opportunity for hybridization between the single-stranded, -terminal nucleotides of the 16S rRNA and the spatially accessible window of mRNA sequence, from which a free energy value can be calculated. Using this algorithm we show that a periodic, energetic pattern of frequency 1/3 is revealed. This periodic signal exists in the majority of coding regions of eubacterial genes, but not in the non-coding regions encoding the 16S and 23S rRNAs. Signal analysis reveals that the population of coding regions of each bacterial species has a mean phase that is correlated in a statistically significant way with species () content. These results suggest that the periodic signal could function as a synchronization signal for the maintenance of reading frame and that codon usage provides a mechanism for manipulation of signal phase.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]  相似文献   

12.
13.
14.
Collapsin response mediator protein 2 (CRMP2) is an intracellular protein that mediates signaling of Semaphorin3A (Sema3A), a repulsive axon guidance molecule. Fyn, a Src-type tyrosine kinase, is involved in the Sema3A signaling. However, the relationship between CRMP2 and Fyn in this signaling pathway is still unknown. In our research, we demonstrated that Fyn phosphorylated CRMP2 at Tyr32 residues in HEK293T cells. Immunohistochemical analysis using a phospho-specific antibody at Tyr32 of CRMP showed that Tyr32-phosphorylated CRMP was abundant in the nervous system, including dorsal root ganglion neurons, the molecular and Purkinje cell layer of adult cerebellum, and hippocampal fimbria. Overexpression of a nonphosphorylated mutant (Tyr32 to Phe32) of CRMP2 in dorsal root ganglion neurons interfered with Sema3A-induced growth cone collapse response. These results suggest that Fyn-dependent phosphorylation of CRMP2 at Tyr32 is involved in Sema3A signaling.Collapsin response mediator proteins (CRMPs)4 have been identified as intracellular proteins that mediate Semaphorin3A (Sema3A) signaling in the nervous system (1). CRMP2 is one of the five members of the CRMP family. CRMPs also mediate signal transduction of NT3, Ephrin, and Reelin (24). CRMPs interact with several intracellular molecules, including tubulin, Numb, kinesin1, and Sra1 (58). CRMPs are involved in axon guidance, axonal elongation, cell migration, synapse maturation, and the generation of neuronal polarity (1, 2, 4, 5).CRMP family proteins are known to be the major phosphoproteins in the developing brain (1, 9). CRMP2 is phosphorylated by several Ser/Thr kinases, such as Rho kinase, cyclin-dependent kinase 5 (Cdk5), and glycogen synthase kinase 3β (GSK3β) (2, 1013). The phosphorylation sites of CRMP2 by these kinases are clustered in the C terminus and have already been identified. Rho kinase phosphorylates CRMP2 at Thr555 (10). Cdk5 phosphorylates CRMP2 at Ser522, and this phosphorylation is essential for sequential phosphorylations by GSK3β at Ser518, Thr514, and Thr509 (2, 1113). These phosphorylations disrupt the interaction of CRMP2 with tubulin or Numb (2, 3, 13). The sequential phosphorylation of CRMP2 by Cdk5 and GSK3β is an essential step in Sema3A signaling (11, 13). Furthermore, the neurofibrillary tangles in the brains of people with Alzheimer disease contain hyperphosphorylated CRMP2 at Thr509, Ser518, and Ser522 (14, 15).CRMPs are also substrates of several tyrosine kinases. The phosphorylation of CRMP2 by Fes/Fps and Fer has been shown to be involved in Sema3A signaling (16, 17). Phosphorylation of CRMP2 at Tyr479 by a Src family tyrosine kinase Yes regulates CXCL12-induced T lymphocyte migration (18). We reported previously that Fyn is involved in Sema3A signaling (19). Fyn associates with PlexinA2, one of the components of the Sema3A receptor complex. Fyn also activates Cdk5 through the phosphorylation at Tyr15 of Cdk5 (19). In dorsal root ganglion (DRG) neurons from fyn-deficient mice, Sema3A-induced growth cone collapse response is attenuated compared with control mice (19). Furthermore, we recently found that Fyn phosphorylates CRMP1 and that this phosphorylation is involved in Reelin signaling (4). Although it has been shown that CRMP2 is involved in Sema3A signaling (1, 11, 13), the relationship between Fyn and CRMP2 in Sema3A signaling and the tyrosine phosphorylation site(s) of CRMPs remain unknown.Here, we show that Fyn phosphorylates CRMP2 at Tyr32. Using a phospho-specific antibody against Tyr32, we determined that the residue is phosphorylated in vivo. A nonphosphorylated mutant CRMP2Y32F inhibits Sema3A-induced growth cone collapse. These results indicate that tyrosine phosphorylation by Fyn at Tyr32 is involved in Sema3A signaling.  相似文献   

15.
16.
Protein ubiquitylation is essential for many events linked to intracellular protein trafficking. Despite the significance of this process, the molecular mechanisms that govern the regulation of ubiquitylation remain largely unknown. Plasma membrane transporters are subjected to tightly regulated endocytosis, and ubiquitylation is a key signal at several stages of the endocytic pathway. The yeast monocarboxylate transporter Jen1 displays glucose-regulated endocytosis. We show here that casein kinase 1-dependent phosphorylation and HECT-ubiquitin ligase Rsp5-dependent ubiquitylation are required for Jen1 endocytosis. Ubiquitylation and endocytosis of Jen1 are induced within minutes in response to glucose addition. Jen1 is modified at the cell surface by oligo-ubiquitylation with ubiquitin-Lys63 linked chain(s), and Jen1-Lys338 is one of the target residues. Ubiquitin-Lys63-linked chain(s) are also required directly or indirectly to sort Jen1 into multivesicular bodies. Jen1 is one of the few examples for which ubiquitin-Lys63-linked chain(s) was shown to be required for correct trafficking at two stages of endocytosis: endocytic internalization and sorting at multivesicular bodies.Ubiquitylation is one of the most prevalent protein post-translational modifications in eukaryotes. In addition to its role in promoting proteasomal degradation of target proteins, ubiquitylation has been shown to regulate multiple processes, including DNA repair, signaling, and intracellular trafficking. Ubiquitylation serves as a key signal mediating the internalization of plasma membrane receptors and transporters, followed by their intracellular transport and subsequent recycling or lysosomal/vacuolar degradation (1, 2). In Saccharomyces cerevisiae, transporters usually display both constitutive and accelerated endocytosis regulated by factors such as excess substrate, changes in nutrient availability, and stress conditions. Ubiquitylation of these cell surface proteins acts as a signal triggering their internalization (1). A single essential E34 ubiquitin ligase, Rsp5, has been implicated in the internalization of most, if not all, endocytosed proteins (3). Rsp5 is the unique member in S. cerevisiae of the HECT (homologous to E6AP COOH terminus)-ubiquitin ligases of the Nedd4/Rsp5 family (4). In a few cases, Rsp5-dependent cell surface ubiquitylation was shown to involve PY-containing adapters that bind to Rsp5 (57). Rsp5-mediated ubiquitylation is also required for sorting into multivesicular bodies (MVBs) of endosomal membrane proteins that come from either the plasma membrane (through endocytosis) or the Golgi (through vacuolar protein sorting (VPS) pathway) (8). Although much progress has been made in elucidating the mechanistic basis of various steps in protein trafficking, the precise requirement for a specific type and length of Ub chains at various stages of the endocytic pathway remains to be addressed.The ubiquitin profile needed for proper internalization has been established for some yeast membrane proteins (1). The α-factor receptor Ste2 was described as undergoing monoubiquitylation on several lysines (multimonoubiquitylation). The a-factor receptor, Ste3p; the general transporter of amino acids, Gap1; the zinc transporter, Ztr1; and the uracil transporter, Fur4, have been shown to be modified by short chains of two to three ubiquitins, each attached to one, two, or more target lysine residues (oligo-ubiquitylation). Among them, Fur4 and Gap1 were the only transporters demonstrated to undergo plasma membrane oligo-ubiquitylation with ubiquitin residues linked via ubiquitin-Lys63 (9, 10). In addition, the two siderophore transporters Arn1 and Sit1 were also shown to undergo Lys63-linked cell surface ubiquitylation (11, 12). Whether these four transporters are representative of a larger class of plasma membrane substrates remains to be determined. Little is known about the type of ubiquitylation involved and/or required for sorting to MVBs. Some MVB cargoes appear to undergo monoubiquitylation (8), whereas Sna3, an MVB cargo of unknown function, undergoes Lys63-linked ubiquitylation (13). Lys63-linked ubiquitin chains were also recently reported to be required, directly or indirectly, for MVB sorting of the siderophore transporter, Sit1, when trafficking through the VPS pathway in the absence of its external substrate (11). In agreement with the possibility that additional membrane-bound proteins might undergo Lys63-linked ubiquitylation, a proteomic study aiming to uncover ubiquitylated yeast proteins showed that Lys63-ubiquitin chains are far more abundant than previously thought (14).The transport of monocarboxylates, such as lactate and pyruvate, as well as ketone bodies across the plasma membrane is essential for the metabolism of cells of various organisms. A family of monocarboxylate transporters has been reported that includes mainly mammalian members (15). In S. cerevisiae, two monocarboxylate-proton symporters have been described, Jen1 and Ady2 (16, 17). These transporters exhibit differences in their mechanisms of regulation and specificity. Jen1 is a lactate-pyruvate-acetate-propionate transporter induced in lactic or pyruvic acid-grown cells (18). Ady2, which accepts acetate, propionate, or formate, is present in cells grown in non-fermentable carbon sources (19). Jen1 has unique regulatory characteristics and has been extensively studied. It was the first secondary porter of S. cerevisiae characterized by heterologous expression in Pichia pastoris at both the cell and the membrane vesicle levels (20). The addition of glucose to lactic acid-grown cells very rapidly triggers loss of Jen1 activity and repression of JEN1 gene expression (21, 22). Newly synthesized Jen1-GFP fusion protein is sorted to the plasma membrane in an active and stable form, and loss of Jen1-GFP activity upon glucose addition is the result of its endocytosis followed by vacuolar degradation (23). Data from large scale analyses based on mass spectrometry approaches led to the detection of two sites of ubiquitylation for Jen1, one located in the N terminus of the protein and the second in the central loop (14), and several sites of phosphorylation in the N terminus, central loop, and C terminus of the protein (14, 24). In the present study, we aimed at further characterizing the internalization step of endocytosis of the transporter Jen1 and the potential role of the phosphorylation and ubiquitylation events required for its correct endocytic trafficking.  相似文献   

17.
Previous studies have shown that protein-protein interactions among splicing factors may play an important role in pre-mRNA splicing. We report here identification and functional characterization of a new splicing factor, Sip1 (SC35-interacting protein 1). Sip1 was initially identified by virtue of its interaction with SC35, a splicing factor of the SR family. Sip1 interacts with not only several SR proteins but also with U1-70K and U2AF65, proteins associated with 5′ and 3′ splice sites, respectively. The predicted Sip1 sequence contains an arginine-serine-rich (RS) domain but does not have any known RNA-binding motifs, indicating that it is not a member of the SR family. Sip1 also contains a region with weak sequence similarity to the Drosophila splicing regulator suppressor of white apricot (SWAP). An essential role for Sip1 in pre-mRNA splicing was suggested by the observation that anti-Sip1 antibodies depleted splicing activity from HeLa nuclear extract. Purified recombinant Sip1 protein, but not other RS domain-containing proteins such as SC35, ASF/SF2, and U2AF65, restored the splicing activity of the Sip1-immunodepleted extract. Addition of U2AF65 protein further enhanced the splicing reconstitution by the Sip1 protein. Deficiency in the formation of both A and B splicing complexes in the Sip1-depleted nuclear extract indicates an important role of Sip1 in spliceosome assembly. Together, these results demonstrate that Sip1 is a novel RS domain-containing protein required for pre-mRNA splicing and that the functional role of Sip1 in splicing is distinct from those of known RS domain-containing splicing factors.Pre-mRNA splicing takes place in spliceosomes, the large RNA-protein complexes containing pre-mRNA, U1, U2, U4/6, and U5 small nuclear ribonucleoprotein particles (snRNPs), and a large number of accessory protein factors (for reviews, see references 21, 22, 37, 44, and 48). It is increasingly clear that the protein factors are important for pre-mRNA splicing and that studies of these factors are essential for further understanding of molecular mechanisms of pre-mRNA splicing.Most mammalian splicing factors have been identified by biochemical fractionation and purification (3, 15, 19, 3136, 45, 6971, 73), by using antibodies recognizing splicing factors (8, 9, 16, 17, 61, 66, 67, 74), and by sequence homology (25, 52, 74).Splicing factors containing arginine-serine-rich (RS) domains have emerged as important players in pre-mRNA splicing. These include members of the SR family, both subunits of U2 auxiliary factor (U2AF), and the U1 snRNP protein U1-70K (for reviews, see references 18, 41, and 59). Drosophila alternative splicing regulators transformer (Tra), transformer 2 (Tra2), and suppressor of white apricot (SWAP) also contain RS domains (20, 40, 42). RS domains in these proteins play important roles in pre-mRNA splicing (7, 71, 75), in nuclear localization of these splicing proteins (23, 40), and in protein-RNA interactions (56, 60, 64). Previous studies by us and others have demonstrated that one mechanism whereby SR proteins function in splicing is to mediate specific protein-protein interactions among spliceosomal components and between general splicing factors and alternative splicing regulators (1, 1a, 6, 10, 27, 63, 74, 77). Such protein-protein interactions may play critical roles in splice site recognition and association (for reviews, see references 4, 18, 37, 41, 47 and 59). Specific interactions among the splicing factors also suggest that it is possible to identify new splicing factors by their interactions with known splicing factors.Here we report identification of a new splicing factor, Sip1, by its interaction with the essential splicing factor SC35. The predicted Sip1 protein sequence contains an RS domain and a region with sequence similarity to the Drosophila splicing regulator, SWAP. We have expressed and purified recombinant Sip1 protein and raised polyclonal antibodies against the recombinant Sip1 protein. The anti-Sip1 antibodies specifically recognize a protein migrating at a molecular mass of approximately 210 kDa in HeLa nuclear extract. The anti-Sip1 antibodies sufficiently deplete Sip1 protein from the nuclear extract, and the Sip1-depleted extract is inactive in pre-mRNA splicing. Addition of recombinant Sip1 protein can partially restore splicing activity to the Sip1-depleted nuclear extract, indicating an essential role of Sip1 in pre-mRNA splicing. Other RS domain-containing proteins, including SC35, ASF/SF2, and U2AF65, cannot substitute for Sip1 in reconstituting splicing activity of the Sip1-depleted nuclear extract. However, addition of U2AF65 further increases splicing activity of Sip1-reconstituted nuclear extract, suggesting that there may be a functional interaction between Sip1 and U2AF65 in nuclear extract.  相似文献   

18.
Mathematical tools developed in the context of Shannon information theory were used to analyze the meaning of the BLOSUM score, which was split into three components termed as the BLOSUM spectrum (or BLOSpectrum). These relate respectively to the sequence convergence (the stochastic similarity of the two protein sequences), to the background frequency divergence (typicality of the amino acid probability distribution in each sequence), and to the target frequency divergence (compliance of the amino acid variations between the two sequences to the protein model implicit in the BLOCKS database). This treatment sharpens the protein sequence comparison, providing a rationale for the biological significance of the obtained score, and helps to identify weakly related sequences. Moreover, the BLOSpectrum can guide the choice of the most appropriate scoring matrix, tailoring it to the evolutionary divergence associated with the two sequences, or indicate if a compositionally adjusted matrix could perform better.[1,2,3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29]  相似文献   

19.
20.
Most individuals infected with human immunodeficiency virus type 1 (HIV-1) initially harbor macrophage-tropic, non-syncytium-inducing (M-tropic, NSI) viruses that may evolve into T-cell-tropic, syncytium-inducing viruses (T-tropic, SI) after several years. The reasons for the more efficient transmission of M-tropic, NSI viruses and the slow evolution of T-tropic, SI viruses remain unclear, although they may be linked to expression of appropriate chemokine coreceptors for virus entry. We have examined plasma viral RNA levels and the extent of CD4+ T-cell depletion in SCID mice reconstituted with human peripheral blood leukocytes following infection with M-tropic, dual-tropic, or T-tropic HIV-1 isolates. The cell tropism was found to determine the course of viremia, with M-tropic viruses producing sustained high viral RNA levels and sparing some CD4+ T cells, dual-tropic viruses producing a transient and lower viral RNA spike and extremely rapid depletion of CD4+ T cells, and T-tropic viruses causing similarly lower viral RNA levels and rapid-intermediate rates of CD4+ T-cell depletion. A single amino acid change in the V3 region of gp120 was sufficient to cause one isolate to switch from M-tropic to dual-tropic and acquire the ability to rapidly deplete all CD4+ T cells.The envelope gene of human immunodeficiency virus type 1 (HIV-1) determines the cell tropism of the virus (11, 32, 47, 62), the use of chemokine receptors as cofactors for viral entry (4, 17), and the ability of the virus to induce syncytia in infected cells (55, 60). Cell tropism is closely linked to but probably not exclusively determined by the ability of different HIV-1 envelopes to bind CD4 and the CC or the CXC chemokine receptors and initiate viral fusion with the target cell. Macrophage-tropic (M-tropic) viruses infect primary cultures of macrophages and CD4+ T cells and use CCR5 as the preferred coreceptor (2, 5, 15, 23, 26, 31). T-cell-tropic (T-tropic) viruses can infect primary cultures of CD4+ T cells and established T-cell lines, but not primary macrophages. T-tropic viruses use CXCR4 as a coreceptor for viral entry (27). Dual-tropic viruses have both of these properties and can use either CCR5 or CXCR4 (and infrequently other chemokine receptors [25]) for viral entry (24, 37, 57). M-tropic viruses are most frequently transmitted during primary infection of humans and persist throughout the duration of the infection (63). Many, but not all, infected individuals show an evolution of virus cell tropism from M-tropic to dual-tropic and finally to T-tropic with increasing time after infection (21, 38, 57). Increases in replicative capacity of viruses from patients with long-term infection have also been noted (22), and the switch to the syncytium-inducing (SI) phenotype in T-tropic or dual-tropic isolates is associated with more rapid disease progression (10, 20, 60). Primary infection with dual-tropic or T-tropic HIV, although infrequent, often leads to rapid disease progression (16, 51). The viral and host factors that determine the higher transmission rate of M-tropic HIV-1 and the slow evolution of dual- or T-tropic variants remain to be elucidated (4).These observations suggest that infection with T-tropic, SI virus isolates in animal model systems with SCID mice grafted with human lymphoid cells or tissue should lead to a rapid course of disease (1, 8, 4446). While some studies in SCID mice grafted with fetal thymus and liver are in agreement with this concept (33, 34), our previous studies with the human peripheral blood leukocyte-SCID (hu-PBL-SCID) mouse model have shown that infection with M-tropic isolates (e.g., SF162) causes more rapid CD4+ T-cell depletion than infection with T-tropic, SI isolates (e.g., SF33), despite similar proviral copy numbers, and that this property mapped to envelope (28, 41, 43). However, the dual-tropic 89.6 isolate (19) caused extremely rapid CD4+ T-cell depletion in infected hu-PBL-SCID mice that was associated with an early and transient increase in HIV-1 plasma viral RNA (29). The relationship between cell tropism of the virus isolate and the pattern of disease in hu-PBL-SCID mice is thus uncertain. We have extended these studies by determining the kinetics of HIV-1 RNA levels in serial plasma samples of hu-PBL-SCID mice infected with primary patient isolates or laboratory stocks that differ in cell tropism and SI properties. The results showed significant differences in the kinetics of HIV-1 replication and CD4+ T-cell depletion that are determined by the cell tropism of the virus isolate.  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