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 46 毫秒
1.
Although K-Ras, Cdc42, and PAK4 signaling are commonly deregulated in cancer, only a few studies have sought to comprehensively examine the spectrum of phosphorylation-mediated signaling downstream of each of these key signaling nodes. In this study, we completed a label-free quantitative analysis of oncogenic K-Ras, activated Cdc42, and PAK4-mediated phosphorylation signaling, and report relative quantitation of 2152 phosphorylated peptides on 1062 proteins. We define the overlap in phosphopeptides regulated by K-Ras, Cdc42, and PAK4, and find that perturbation of these signaling components affects phosphoproteins associated with microtubule depolymerization, cytoskeletal organization, and the cell cycle. These findings provide a resource for future studies to characterize novel targets of oncogenic K-Ras signaling and validate biomarkers of PAK4 inhibition.The Ras oncoproteins are small monomeric GTPases that transduce mitogenic signals from cell surface receptor tyrosine kinases (RTKs) to intracellular serine/threonine kinases. Approximately thirty percent of human tumors harbor a somatic gain-of-function mutation in one of three RAS genes, resulting in the constitutive activation of Ras signaling and the aberrant hyperactivation of growth-promoting effector pathways (1). Designing therapeutic agents that directly target Ras has been challenging (2, 3), and thus clinical development efforts have focused on targeting effector pathways downstream of Ras. The Raf-MEK-ERK and PI3K-Akt effector pathways have been extensively studied and several small molecule inhibitors targeting these pathways are currently under clinical evaluation (4, 5). However, biochemical studies and mouse models indicate that several additional effector pathways are essential for Ras-driven transformation and tumorigenesis (611). Hence, a comprehensive characterization of these effector pathways may reveal additional druggable targets.The Rho GTPase Cdc42 lies downstream of Ras (1214) and regulates many cellular processes that are commonly perturbed in cancer, including migration, polarization, and proliferation (15) (Fig. 1A). Importantly, Cdc42 is overexpressed in several types of human cancer (1620) and is required for Ras-driven cellular transformation (13, 21, 22). Recent studies show that genetic ablation of Cdc42 impairs Ras-driven tumorigenesis (13), indicating the potential of Cdc42 and its effectors as drug targets in Ras mutant tumors.Open in a separate windowFig. 1.Experimental workflow. A, K-Ras is a small GTPase that regulates the activity of a variety of downstream proteins including the Rho GTPase Cdc42. The PAK4 serine/threonine kinase is a direct effector of Cdc42 and regulates actin reorganization, microtubule stability, and cell polarity. B, To measure large-scale phosphorylation changes induced by constitutive K-Ras or Cdc42 signaling or PAK4 ablation, the quantitative label-free PTMscan® approach was employed (Cell Signaling Technology). Briefly, for each condition extracted proteins were digested with trypsin and separated from non-peptide material by solid-phase extraction with Sep-Pak C18 cartridges. Three phosphorylation motif antibodies were used serially to isolate phosphorylated peptides in independent immunoaffinity purifications (CDK substrate motif [K R]-pS-P-X-[K R], CK substrate motif pT-[D E]-X-[D E], PKD substrate motif l-X-R-X-X-p[S T]). The samples were run in duplicate and tandem mass spectra were collected with an LTQ-Orbitrap hybrid mass spectrometer. pLPC is an empty vector control.In particular, the p21-activated kinases (PAKs) are Cdc42 effectors that have generated significant interest (23, 24), as they are central components of key oncogenic signaling pathways and regulate cytoskeletal organization, cell migration, and nuclear signaling (25). The PAK family is comprised of six members and is subdivided into two groups (Groups I and II) based on sequence and structural homology. Group I PAKs (PAK1–3) are relatively well characterized, however, much less is known regarding the function and regulation of Group II PAKs (PAK4–6). The kinase domains of Group I and II PAKs share only about 50% identity, suggesting the two groups may recognize distinct substrates and govern unique cellular processes (26).The Group II PAK family member PAK4 is of particular interest as it is overexpressed or genetically amplified in several lung, colon, prostate, pancreas, and breast tumor cell lines and samples (2630). Furthermore, functional studies have implicated PAK4 in cell transformation, cell invasion, and migration (27, 31). Xenograft studies in athymic mice show an important role for PAK4 in mediating Cdc42- or K-Ras-driven tumor formation, highlighting a critical role for Pak4 downstream of these GTPases (32). Given its roles in transformation, tumorigenesis, and oncogenic signaling, there is significant interest in targeting PAK4 therapeutically (23). PAK4 binds and phosphorylates several proteins involved in cytoskeletal organization and apoptosis, including Lim domain kinase 1 (LIMK1) (33), guanine nucleotide exchange factor-H1 (GEF-H1) (34), Raf-1 (35), and Bad (36). However, the Group I PAK family member PAK1 also phosphorylates several of these PAK4 targets (37). Thus, there remains a need to identify robust and selective pharmacodynamic biomarkers for PAK4 inhibition.Despite the importance of PAK4 and its upstream regulators in cancer development, few studies have sought to comprehensively characterize the spectrum of K-Ras, Cdc42, or PAK4 mediated phosphorylation signaling (3739). Recent developments in mass spectrometry allow the in-depth identification and quantitation of thousands of phosphorylation sites (4043). The majority of large-scale efforts have aimed to identify the basal phosphoproteomes of different species (44, 45) or tissues (46) to characterize global steady-state phosphorylation. However, this methodology can also be applied to quantify perturbed phosphorylation regulation in cancer signaling pathways (40, 4749), and has the potential to reveal novel biomarkers of oncogenic signaling.In this study, we completed a label-free quantitative analysis of K-Ras, Cdc42, and PAK4 phosphorylation signaling using the PTMScan® method, which has proven as robust and reproducible quantitation technology (50, 51). We quantified phosphorylation levels in wild-type and PAK4 knockout NIH3T3 cells expressing oncogenic K-Ras, activated Cdc42, or an empty vector control to elucidate the molecular pathways and functions modulated by these key signaling proteins. We report relative quantitation of 2152 phosphorylated peptides on 1062 proteins among the different conditions, and find that many of the regulated phosphoproteins are associated with microtubule depolymerization, cytoskeletal organization, and the cell cycle. To our knowledge, our study is the first to examine the overlap among signaling networks regulated by K-Ras, Cdc42, and PAK4, and provides a resource for future studies to further interrogate the perturbation of this signaling pathway.  相似文献   

2.
Complex I (EC 1.6.99.3) of the bacterium Escherichia coli is considered to be the minimal form of the type I NADH dehydrogenase, the first enzyme complex in the respiratory chain. Because of its small size and relative simplicity, the E. coli enzyme has become a model used to identify and characterize the mechanism(s) by which cells regulate the synthesis and assembly of this large respiratory complex. To begin dissecting the processes by which E. coli cells regulate the expression of nuo and the assembly of complex I, we undertook a genetic analysis of the nuo locus, which encodes the 14 Nuo subunits comprising E. coli complex I. Here we present the results of studies, performed on an isogenic collection of nuo mutants, that focus on the physiological, biochemical, and molecular consequences caused by the lack of or defects in several Nuo subunits. In particular, we present evidence that NuoG, a peripheral subunit, is essential for complex I function and that it plays a role in the regulation of nuo expression and/or the assembly of complex I.

Complex I (NADH:ubiquinone oxidoreductase; EC 1.6.99.3), a type I NADH dehydrogenase that couples the oxidation of NADH to the generation of a proton motive force, is the first enzyme complex of the respiratory chain (2, 35, 47). The Escherichia coli enzyme, considered to be the minimal form of complex I, consists of 14 subunits instead of the 40 to 50 subunits associated with the homologous eukaryotic mitochondrial enzyme (17, 29, 30, 4850). E. coli also possesses a second NADH dehydrogenase, NDH-II, which does not generate a proton motive force (31). E. coli complex I resembles eukaryotic complex I in many ways (16, 17, 30, 49): it performs the same enzymatic reaction and is sensitive to a number of the same inhibitors, it consists of subunits homologous to those found in all proton-translocating NADH:ubiquinone oxidoreductases studied thus far, it is comprised of a large number of subunits relative to the number that comprise other respiratory enzymes, and it contains flavin mononucleotide and FeS center prosthetic groups. Additionally, it possesses an L-shaped topology (14, 22) like that of its Neurospora crassa homolog (27), and it consists of distinct fragments or subcomplexes. Whereas eukaryotic complex I can be dissected into a peripheral arm and a membrane arm, the E. coli enzyme consists of three subcomplexes referred to as the peripheral, connecting, and membrane fragments (29) (Fig. (Fig.1A).1A). The subunit composition of these three fragments correlates approximately with the organization of the 14 structural genes (nuoA to nuoN) (49) of the nuo (for NADH:ubiquinone oxidoreductase) locus (Fig. (Fig.1B),1B), an organization that is conserved in several other bacteria, including Salmonella typhimurium (3), Paracoccus denitrificans (53), Rhodobacter capsulatus (12), and Thermus thermophilus (54). The 5′ half of the locus contains a promoter (nuoP), previously identified and located upstream of nuoA (8, 49), and the majority of genes that encode subunits homologous to the nucleus-encoded subunits of eukaryotic complex I and to subunits of the Alcaligenes eutrophus NAD-reducing hydrogenase (17, 29, 30, 49). In contrast, the 3′ half contains the majority of the genes that encode subunits homologous to the mitochondrion-encoded subunits of eukaryotic complex I and to subunits of the E. coli formate-hydrogen lyase complex (17, 29, 30, 49). Whereas the nuclear homologs NuoE, NuoF, and NuoG constitute the peripheral fragment (also referred to as the NADH dehydrogenase fragment [NDF]), the nuclear homologs NuoB, NuoC, NuoD, and NuoI constitute the connecting fragment. The mitochondrial homologs NuoA, NuoH, NuoJ, NuoK, NuoL, NuoM, and NuoN constitute the membrane fragment (29). E. coli complex I likely evolved by fusion of preexisting protein assemblies constituting modules for electron transfer and proton translocation (1719, 30). Open in a separate windowFIG. 1Schematic of E. coli complex I and the nuo locus. Adapted with permission of the publisher (17, 29, 30, 49). (A) E. coli complex I is comprised of three distinct fragments: the peripheral (light gray), connecting (white), and membrane (dark gray) fragments (17, 29). The peripheral fragment (NDF) is comprised of the nuclear homologs NuoE, -F, and -G and exhibits NADH dehydrogenase activity that oxidizes NADH to NAD+; the connecting fragment is comprised of the nuclear homologs NuoB, -C, -D, and -I; and the membrane fragment is comprised of the mitochondrial homologs NuoA, -H, and -J to -N and catalyzes ubiquinone (Q) to its reduced form (QH2). FMN, flavin mononucleotide. (B) The E. coli nuo locus encodes the 14 Nuo subunits that constitute complex I. The 5′ half of the locus contains a previously identified promoter (nuoP) and the majority of genes that encode the peripheral and connecting subunits (light gray and white, respectively). The 3′ half of the locus contains the majority of the genes encoding the membrane subunits (dark gray). The 3′ end of nuoG encodes a C-Terminal region (CTR) of the NuoG subunit (hatched).Because of its smaller size and relative simplicity, researchers recently have begun to utilize complex I of E. coli, and that of its close relative S. typhimurium, to identify and characterize the mechanism(s) by which cells regulate the synthesis and assembly of this large respiratory complex (3, 8, 46) and to investigate the diverse physiological consequences caused by defects in this enzyme (4, 6, 10, 40, 59). Such defects affect the ability of cells to perform chemotaxis (40), to grow on certain carbon sources (4, 6, 10, 40, 57), to survive stationary phase (59), to perform energy-dependent proteolysis (4), to regulate the expression of at least one gene (32), and to maintain virulence (5).To begin dissecting the processes by which E. coli cells regulate the expression of nuo and the assembly of complex I, we undertook a genetic analysis of the nuo locus. Here, we present the results of studies, performed on an isogenic collection of nuo mutants, that focus on the physiological, biochemical, and molecular consequences caused by the lack of or defects in several Nuo subunits. In particular, we present evidence that NuoG, a peripheral subunit, is essential for complex I function and that it plays a role in the regulation of nuo expression and/or the assembly of complex I.  相似文献   

3.
4.
5.
6.
7.
8.
Tandem repeat (TR) regions are common in yeast adhesins, but their structures are unknown, and their activities are poorly understood. TR regions in Candida albicans Als proteins are conserved glycosylated 36-residue sequences with cell-cell aggregation activity (J. M. Rauceo, R. De Armond, H. Otoo, P. C. Kahn, S. A. Klotz, N. K. Gaur, and P. N. Lipke, Eukaryot. Cell 5:1664–1673, 2006). Ab initio modeling with either Rosetta or LINUS generated consistent structures of three-stranded antiparallel β-sheet domains, whereas randomly shuffled sequences with the same composition generated various structures with consistently higher energies. O- and N-glycosylation patterns showed that each TR domain had exposed hydrophobic surfaces surrounded by glycosylation sites. These structures are consistent with domain dimensions and stability measurements by atomic force microscopy (D. Alsteen, V. Dupres, S. A. Klotz, N. K. Gaur, P. N. Lipke, and Y. F. Dufrene, ACS Nano 3:1677–1682, 2009) and with circular dichroism determination of secondary structure and thermal stability. Functional assays showed that the hydrophobic surfaces of TR domains supported binding to polystyrene surfaces and other TR domains, leading to nonsaturable homophilic binding. The domain structures are like “classic” subunit interaction surfaces and can explain previously observed patterns of promiscuous interactions between TR domains in any Als proteins or between TR domains and surfaces of other proteins. Together, the modeling techniques and the supporting data lead to an approach that relates structure and function in many kinds of repeat domains in fungal adhesins.Yeast adhesins are a diverse set of cell adhesion proteins that mediate adhesion to host cells, environmental substrates, other fungi, and coinfecting bacteria (6, 8, 20, 21, 23, 29). The adhesins share common features, including compact N-terminal domains similar to Ig or lectin domains, Thr-rich midpieces, often in tandem repeats, and long highly glycosylated Ser/Thr-rich C-terminal regions that extend the functional domains out from the cell surface. No structures for the Thr-rich midpieces are known, but they can mediate aggregation of fungal cells (33, 35, 47). The prevalence and conservation of such repeats argue that they are functionally important, despite limited data on their structure and function.In Candida albicans, the Als adhesins are homologous proteins, products of 8 loci that encode numerous alleles of cell surface adhesins (16). In each mature Als protein, there are, from the N terminus, three tandem Ig-like domains, a β-sheet-rich conserved 127-residue amyloid-forming T region, a variable number of 36-residue tandem repeats (TRs), and a highly glycosylated stalk region that extends the N-terminal domains away from the cell surface (Fig. 1) (16, 33, 41). The C termini of these and other wall-associated adhesins are covalently cross-linked into the cell wall through transglycosylation of a modified glycosylphosphatidylinositol (GPI) anchor (18, 25). This modular design, including tandem repeats, is typical of fungal adhesins (8).Open in a separate windowFig. 1.Schematic diagram of the sequence of Als5p. The regions are named above, and the number of amino acid residues in each region is shown below. The modeled sequences are in the TR region.The Als protein Ig-like region, T region, and TR region all have protein-protein interaction activities (26, 33, 35). The Ig-like regions can interact with diverse mammalian proteins, presumably in a way analogous to antibody-antigen binding, as has been shown in the homologous protein α-agglutinin from Saccharomyces cerevisiae (8, 24, 26, 35). The T regions interact through formation of amyloid-like structures both in vivo and in vitro (33, 34a, 36). An insight into the function of the tandem repeats followed from observations that Als proteins initiate and maintain cell-to-cell aggregations, either spontaneously (“autoaggregation”) or following adhesion to a bead-bound defined ligand (10, 11, 36). Aggregation is more extensive for Als proteins with more tandem repeats (26, 35). This result suggested that the tandem repeats are uniquely structured to facilitate or mediate the aggregative function. Circular dichroism spectroscopy of the TR region of Als5p shows a β-sheet-rich structure in the soluble protein (35).In support of their direct involvement in aggregation, the repeat region of the C. albicans adhesin Als5p mediates cell-cell aggregation in the absence of the Ig-like and T domains (35). Moreover, the repeats can also potentiate binding of Als5p to fibronectin (35). Thus, the TR domains mediate cellular aggregation and increased binding to fibronectin. In addition, TR domains and their amino acid sequences are highly conserved across several Candida species (3). These properties need to be explained by their three-dimensional structure.Because there are no homologous structures known, we modeled by two independent ab initio methods. Rosetta assembles structures by combining short peptide structures extracted from the protein structural database PDB (38), then combines structures in a Monte Carlo approach, and assesses energetics of assembled structures. Rosetta has recently been shown to generate accurate models for protein-sized domains (40). We also predicted structures with LINUS, which generates randomized structures and rapidly estimates energetics to choose low-energy models (45). The models were supported by structural analyses with atomic force microscopy and circular dichroism spectroscopy. Functional assays showed that the TR domains can mediate binding activities predicted from the calculated structures.  相似文献   

9.
10.
11.
Oxidative modifications of protein tyrosines have been implicated in multiple human diseases. Among these modifications, elevations in levels of 3,4-dihydroxyphenylalanine (DOPA), a major product of hydroxyl radical addition to tyrosine, has been observed in a number of pathologies. Here we report the first proteome survey of endogenous site-specific modifications, i.e. DOPA and its further oxidation product dopaquinone in mouse brain and heart tissues. Results from LC-MS/MS analyses included 50 and 14 DOPA-modified tyrosine sites identified from brain and heart, respectively, whereas only a few nitrotyrosine-containing peptides, a more commonly studied marker of oxidative stress, were detectable, suggesting the much higher abundance for DOPA modification as compared with tyrosine nitration. Moreover, 20 and 12 dopaquinone-modified peptides were observed from brain and heart, respectively; nearly one-fourth of these peptides were also observed with DOPA modification on the same sites. For both tissues, these modifications are preferentially found in mitochondrial proteins with metal binding properties, consistent with metal-catalyzed hydroxyl radical formation from mitochondrial superoxide and hydrogen peroxide. These modifications also link to a number of mitochondrially associated and other signaling pathways. Furthermore, many of the modification sites were common sites of previously reported tyrosine phosphorylation, suggesting potential disruption of signaling pathways. Collectively, the results suggest that these modifications are linked with mitochondrially derived oxidative stress and may serve as sensitive markers for disease pathologies.Generation of reactive oxygen species (ROS)1 and reactive nitrogen species is a normal consequence of aerobic metabolism that, in excess, results in oxidative stress that further leads to oxidative modification of proteins, lipids, and DNA, events that may lead to altered cellular function and even cell death (1, 2). Chronic oxidative stress is well recognized as having a central role in disease and is responsible for both direct alteration of biomolecular structure-function and compensatory changes in cellular processes (14). It is increasingly recognized that oxidative modifications of proteins can serve as potential biomarkers indicative of the physiological states and changes that occur during disease progression. Thus, the ability to quantitatively measure specific protein oxidation products has the potential to provide the means to monitor the physiological state of a tissue or organism, in particular any progression toward pathology. Given Parkinson disease (PD) as an example, a number of oxidative modifications on proteins pertinent to PD have been identified, further supporting the potential importance of oxidative modifications to disease pathogenesis (5).Many oxidative modifications on specific amino acid residues, such as protein carbonylation (6), cysteine S-nitrosylation (79), cysteine oxidation to sulfinic or sulfonic acid (1012), methionine oxidation (13, 14), and tyrosine nitration (1521) within complex protein mixtures, have been detected by MS-based proteomics; however, their low abundance levels within complex proteomes often hinder confident identification of these potentially significant modifications (22). For example, tyrosine nitration is a well studied post-translational modification mediated by peroxynitrite (ONOO) or nitrogen dioxide (·NO2), which commonly occur in cells during oxidative stress and inflammation; however, only a small number of nitrotyrosine proteins have been identified from a given proteome sample because of insufficient analytical sensitivity and the chance of incorrect peptide assignments (19, 23). With recent advances in high resolution MS that provide high mass measurement accuracy, the ability to confidently identify modified peptides has been significantly enhanced (24).Hydroxyl radical (HO·) is one of the most reactive and major species generated under aerobic conditions in biological systems (1, 25, 26). Among several HO·-mediated oxidative modifications, the protein tyrosine modification 3,4-dihydroxyphenylalanine (DOPA) has been reported as a major product and index of HO· attack on tyrosine residues in proteins (Fig. 1) (27, 28). DOPA is also formed on protein tyrosine residues via controlled enzymatic pathways through enzymes such as tyrosinase or tyrosine hydroxylase (28). Once formed, protein-bound DOPA has the potential to initiate further oxidative reactions through binding and reducing transition metals or through redox cycling between catechol and quinone (dopaquinone) forms (29, 30). Recent studies have suggested that protein-bound DOPA is involved in triggering antioxidant defenses (30) and mediating oxidative damage to DNA (31). Moreover, elevated levels of protein-bound DOPA have been reported in several diseases, including atherosclerosis, cataracts, and myocardial disease, and in PD patients undergoing levodopa therapy (26, 3236). However, the specific DOPA-modified proteins, which could provide mechanistic knowledge of the progression of these diseases, have not been identified (27, 28). The ability to identify site-specific protein modifications should lead to a better understanding of the role of DOPA modification in disease pathologies as well as new molecular signatures or therapeutic targets for diseases.Open in a separate windowFig. 1.DOPA and dopaquinone formation from tyrosine.Therefore, in this study, we demonstrate the ability to identify site-specific DOPA and dopaquinone (DQ) modifications on protein tyrosine residues in normal mouse brain and heart tissues and their relative stoichiometries that are present in vivo under non-stressed conditions. Such endogenous protein modifications were detected using LC-MS/MS. The results from this global proteomics survey suggests that HO· in tissues under normal conditions is generated largely from the mitochondria and metal-binding proteins where the resulting DOPA/DQ modifications have the potential to disrupt mitochondrial respiration as well as alter tyrosine phosphorylation signaling pathways such as 14-3-3-mediated signaling in brain tissue.  相似文献   

12.
An epoxide hydrolase from Rhodococcus erythropolis DCL14 catalyzes the hydrolysis of limonene-1,2-epoxide to limonene-1,2-diol. The enzyme is induced when R. erythropolis is grown on monoterpenes, reflecting its role in the limonene degradation pathway of this microorganism. Limonene-1,2-epoxide hydrolase was purified to homogeneity. It is a monomeric cytoplasmic enzyme of 17 kDa, and its N-terminal amino acid sequence was determined. No cofactor was required for activity of this colorless enzyme. Maximal enzyme activity was measured at pH 7 and 50°C. None of the tested inhibitors or metal ions inhibited limonene-1,2-epoxide hydrolase activity. Limonene-1,2-epoxide hydrolase has a narrow substrate range. Of the compounds tested, only limonene-1,2-epoxide, 1-methylcyclohexene oxide, cyclohexene oxide, and indene oxide were substrates. This report shows that limonene-1,2-epoxide hydrolase belongs to a new class of epoxide hydrolases based on (i) its low molecular mass, (ii) the absence of any significant homology between the partial amino acid sequence of limonene-1,2-epoxide hydrolase and amino acid sequences of known epoxide hydrolases, (iii) its pH profile, and (iv) the inability of 2-bromo-4′-nitroacetophenone, diethylpyrocarbonate, 4-fluorochalcone oxide, and 1,10-phenanthroline to inhibit limonene-1,2-epoxide hydrolase activity.Epoxides are highly reactive compounds which readily react with numerous biological compounds, including proteins and nucleic acids. Consequently, epoxides are cytotoxic, mutagenic, and potentially carcinogenic, and there is considerable interest in biological degradation mechanisms for these compounds.In bacteria, epoxides are formed during the metabolism of alkenes (23) and halohydrins (15, 26, 34, 49). Enzymes belonging to a large number of enzyme classes, including dehydrogenases (17), lyases (21), carboxylases (1, 43), glutathione S-transferases (6, 8), isomerases (24), and hydrolases (7, 19, 44), are involved in the microbial degradation of epoxides.Epoxide hydrolases are enzymes catalyzing the addition of water to epoxides forming the corresponding diol. This group of enzymes has been extensively studied in mammals, while only limited information is available on bacterial epoxide hydrolases. Three functions for epoxide hydrolases are recognized (42). In bacteria, epoxide hydrolases are involved in the degradation of several hydrocarbons, including 1,3-dihalo-2-propanol (34), 2,3-dihalo-1-propanol (15, 26), epichlorohydrin (46), propylene oxide (16), 9,10-epoxy fatty acids (30, 36), trans-2,3-epoxysuccinate (2), and cyclohexene oxide (14). Other epoxide hydrolases, such as microsomal and cytosolic epoxide hydrolase from mammals (for reviews, see references 4, 8, and 44), are involved in the detoxification of epoxides formed due to the action of P-450-dependent monooxygenases (8). Epoxide hydrolases are also involved in biosynthesis of hormones, such as leukotrienes and juvenile hormone (40, 45), and plant cuticular elements (11). Remarkably, the bacterial and eukaryotic epoxide hydrolases described so far form a homogeneous group of enzymes belonging to the α/β-hydrolase fold superfamily (10, 38).Rhodococcus erythropolis DCL14, a gram-positive bacterium, is able to grow on both (+)- and (−)-limonene as the sole source of carbon and energy (47). Cells grown on limonene contained a novel epoxide hydrolase that does not belong to the α/β-hydrolase fold superfamily. This limonene-1,2-epoxide hydrolase converts limonene-1,2-epoxide to limonene-1,2-diol (p-menth-8-ene-1,2-diol [Fig. 1]). In this report, we describe the purification and characterization of this enzyme and show that limonene-1,2-epoxide hydrolase belongs to a novel class of epoxide hydrolases. Open in a separate windowFIG. 1Reaction catalyzed by limonene-1,2-epoxide hydrolase.  相似文献   

13.
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   

14.
15.
IgA nephropathy (IgAN) is the most common primary glomerulonephritis in the world. Aberrantly glycosylated IgA1, with galactose (Gal)-deficient hinge region (HR) O-glycans, plays a pivotal role in the pathogenesis of the disease. It is not known whether the glycosylation defect occurs randomly or preferentially at specific sites. We have described the utility of activated ion-electron capture dissociation (AI-ECD) mass spectrometric analysis of IgA1 O-glycosylation. However, locating and characterizing the entire range of O-glycan attachment sites are analytically challenging due to the clustered serine and threonine residues in the HR of IgA1 heavy chain. To address this problem, we analyzed all glycoforms of the HR glycopeptides of a Gal-deficient IgA1 myeloma protein, mimicking the aberrant IgA1 in patients with IgAN, by use of a combination of IgA-specific proteases + trypsin and AI-ECD Fourier transform ion cyclotron resonance (FT-ICR) tandem mass spectrometry (MS/MS). The IgA-specific proteases provided a variety of IgA1 HR fragments that allowed unambiguous localization of all O-glycosylation sites in the six most abundant glycoforms, including the sites deficient in Gal. Additionally, this protocol was adapted for on-line liquid chromatography (LC)-AI-ECD MS/MS and LC-electron transfer dissociation MS/MS analysis. Our results thus represent a new clinically relevant approach that requires ECD/electron transfer dissociation-type fragmentation to define the molecular events leading to pathogenesis of a chronic kidney disease. Furthermore, this work offers generally applicable principles for the analysis of clustered sites of O-glycosylation.Glycosylation is one of the most common post-translational modifications of proteins. It is estimated that over half of mammalian proteins are glycosylated. Patients with several autoimmune disorders, chronic inflammatory diseases, and some infectious diseases exhibit abnormal glycosylation of serum immunoglobulins and other glycoproteins (15). The biological functions of these modifications in health and disease have become a significant area of interest in biomedical research (6). A subset of these glycoproteins has clustered sites of O-glycosylation with serine- and threonine-rich stretches within the amino acid sequence. Mucins, such as membrane-associated MUC1, are perhaps the best known family of proteins that are heavily O-glycosylated. Their altered expression and aberrant glycosylation have made them potential targets as biomarkers for early detection of cancer (7). Immunoglobulin A1 (IgA1)1 contains both O- and N-glycans (Fig. 1). Aberrant O-glycosylation of IgA1 is involved in the pathogenesis of IgA nephropathy (IgAN) and the closely related Henoch-Schönlein purpura nephritis (1, 8). Interestingly, the aberrantly glycosylated molecules, IgA1 in IgAN and MUC1 in cancer, are recognized by the immune system as neoepitopes as evidenced by formation of specific antibodies (911). Mucin-like bacterial surface proteins exhibit similar properties: the molecules have clustered bacterial O-glycans that mediate cellular adhesion, and blocking antibodies target these glycan-containing epitopes (12).Open in a separate windowFig. 1.IgA1 structural elements. IgA1 has N-linked glycans (filled circles) and O-linked glycans (open circles). The O-glycosylated sites are in the HR between the first and second constant region domains of the heavy chains. The HR is a Pro-rich segment with nine possible sites of O-glycan attachment. Underlined serine and threonine residues are usually glycosylated (31). Arrows show cleavage sites of trypsin and IgA-specific proteases.An O-glycosylated protein from a single source contains a population of variably O-glycosylated isoforms that show a distinct distribution of microheterogeneity of the O-glycan chains in terms of number, sites of attachment, and composition. Characterizing these clustered sites and understanding how the distributions change under different biological conditions or disease states are an analytical challenge. Enzymatic or chemical release of O-glycans is not selective. The heterogeneity, composition, and quantitative aspects of different O-glycan chains can be assessed and quantified by gas chromatographic and/or mass spectrometric techniques. However, the site-specific information and context of location and composition of adjacent chains are lost. Carbohydrate-specific lectin analysis of O-glycoproteins can provide information on glycan composition and comparative differences between samples, such as those from healthy controls and patients with various disease states. We have successfully demonstrated this in the analysis of IgA1 O-glycans from patients with IgAN versus healthy controls and disease controls (1315). This included proximal assessment of sites with galactose (Gal)-deficient O-glycans after digests with IgA-specific proteases (8). Several studies have demonstrated the value of mass spectrometry (MS) in identifying Gal-deficient IgA1 in patients with IgAN (1621), including our work that demonstrated the first direct localization of native sites of O-glycan chains in the hinge region (HR) of IgA1 by use of electron capture dissociation (ECD) (20, 22). ECD and the more recently developed electron transfer dissociation (ETD) have been used to identify sites of O-glycosylation on a variety of proteins (2326). This includes the analysis of sites of O-glycosylation by on-line LC-ECD/ETD MS/MS methods (23, 26, 27).IgAN is the most common primary glomerulonephritis worldwide (28) with about 20–40% of patients developing end stage renal failure. It is characterized by mesangial deposits of IgA1-containing immune complexes (28). The distinctive O-glycan chains of IgA1 molecules play a pivotal role in the pathogenesis of IgAN (1, 10, 1416, 29, 30). IgA1 contains an HR between the first and second heavy chain constant region domains with a high content of Ser, Thr, and Pro. This segment usually has three to five O-glycan chains per HR (31) (see Fig. 1). Aberrantly glycosylated IgA1, deficient in Gal in some of the O-glycans in the HR, in serum is rare in healthy individuals but is present at elevated levels in IgAN patients (13, 15). This distinctive IgA1 is in circulating immune complexes (8, 10, 15) and in the glomerular deposits of IgAN patients (16, 29). The absence of Gal apparently leads to the exposure of neoepitopes, including terminal and sialylated N-acetylgalactosamine (GalNAc) residues (9, 10). These epitopes are recognized by naturally occurring anti-glycan IgG or IgA1 antibodies and, consequently, circulating immune complexes are formed (9, 10, 15) that can deposit in the glomerular mesangia. To identify the pathogenic forms of IgA1, a thorough analysis of O-glycan microheterogeneity, including identification of the attachment sites, will be required.In this work, we demonstrate the complete analysis of O-glycoform microheterogeneity and site localization of the glycoforms in a naturally Gal-deficient IgA1 (Ale) myeloma protein that mimics the nephritogenic IgA1 in patients with IgAN (8, 9). Reversed phase (RP) LC FT-ICR MS successfully identified 10 distinct IgA1 HR fragments representing >99% of total IgA1. AI-ECD of the six most abundant IgA1 HR glycoforms (>95% of total IgA1) was accomplished with three distinct IgA-specific protease + trypsin digestions, identifying sites of Gal deficiency across four distinct IgA1 O-glycoforms. Based on the success of the ECD fragmentation of these IgA1 HR fragments, we adapted the analysis for on-line LC-MS/MS methods for both ECD and ETD. The variety of IgA1 HR proteolytic fragments provides a practical set of guidelines for the ECD/ETD analysis of clustered sites of O-glycosylation on this and other proteins. These results also provide insight into the order of attachment of the O-glycans in the IgA1 HR.  相似文献   

16.
17.
18.
19.
Linear genome stability requires specialized telomere replication and protection mechanisms. A common solution to this problem in non-eukaryotes is the formation of hairpin telomeres by telomere resolvases (also known as protelomerases). These enzymes perform a two-step transesterification on replication intermediates to generate hairpin telomeres using an active site similar to that of tyrosine recombinases and type IB topoisomerases. Unlike phage telomere resolvases, the telomere resolvase from the Lyme disease pathogen Borrelia burgdorferi (ResT) is a permissive enzyme that resolves several types of telomere in vitro. However, the ResT region and residues mediating permissive substrate usage have not been identified. The relapsing fever Borrelia hermsii ResT exhibits a more restricted substrate usage pattern than B. burgdorferi ResT and cannot efficiently resolve a Type 2 telomere. In this study, we determined that all relapsing fever ResTs process Type 2 telomeres inefficiently. Using a library of chimeric and mutant B. hermsii/B. burgdorferi ResTs, we mapped the determinants in B. burgdorferi ResT conferring the ability to resolve multiple Type 2 telomeres. Type 2 telomere resolution was dependent on a single proline in the ResT catalytic region that was conserved in all Lyme disease but not relapsing fever ResTs and that is part of a 2-amino acid insertion absent from phage telomere resolvase sequences. The identification of a permissive substrate usage determinant explains the ability of B. burgdorferi ResT to process the 19 unique telomeres found in its segmented genome and will aid further studies on the structure and function of this essential enzyme.Replication and protection of telomeric DNA are required to ensure the genomic stability of all organisms with linear replicons. Until quite recently, it was assumed that linearity is a property confined to the replicons of eukaryotes and certain primarily eukaryotic viruses. However, a growing body of evidence indicates that linear DNA is also found in a broad range of bacteriophages (16) and in bacteria themselves (710), including the Borrelia species that cause Lyme disease and relapsing fever (11, 12). A common solution to the end replication and protection problem in non-eukaryotes is the covalent sealing of DNA ends in the form of hairpins (2, 46, 10, 11, 1316). Hairpin DNA is not recognized as a double-strand break, and continuous synthesis of DNA around the hairpin loop abolishes the end replication problem. However, mother and daughter replicons are covalently linked at the junction of their telomeres following DNA replication; separation of the two replicons and formation of new hairpin telomeres require a DNA breakage and reunion process referred to as telomere resolution (17, 18).Resolution of the linear chromosome and plasmids in Borrelia species and of the linear plasmid prophages from Escherichia coli, Yersinia enterocolitica, and Klebsiella oxytoca is performed by telomere resolvases (also referred to as protelomerases) (5, 1921). A growing number of candidate telomere resolvases have been identified in the genomes of eukaryotic viruses, phages, and bacteria (22, 23). Telomere resolvases are DNA cleavage and rejoining enzymes related to tyrosine recombinases and type 1B topoisomerases (19, 21, 22, 24, 25). Telomere resolvase catalyzes a two-step transesterification reaction in which staggered cuts are introduced 6 bp apart on either side of the axis of symmetry in the replicated telomere substrate (5, 19, 21, 24). Cleavage is accompanied by the formation of a 3′-phosphotyrosyl protein-DNA linkage. Subsequent nucleophilic attack on opposing strands by the free 5′-OH groups in the nicked substrate creates covalently closed hairpin telomeres. A recent crystal structure of the Klebsiella phage telomere resolvase (TelK) in complex with its substrate identified the residues involved in catalysis (25); all but one of these residues are conserved in all telomere resolvases (22), implying that the basic catalytic mechanism underlying telomere resolution is conserved. However, telomere resolvase sequences vary substantially outside of the central catalytic region (25, 26), and the enzymes characterized to date demonstrate important differences in substrate usage that likely reflect functionally distinct mechanisms of substrate interaction.The Borrelia burgdorferi telomere resolvase, ResT, appears to be particularly divergent. It is substantially smaller than phage telomere resolvases, and unlike its phage counterparts (5, 20, 21), it cannot efficiently resolve negatively supercoiled DNA (19, 27), presumably reflecting differences in the substrates resolved by phage and Borrelia telomere resolvases in vivo. On the other hand, B. burgdorferi ResT can fuse hairpin telomeres in a reversal of the resolution reaction (28), a function that is not shared with the phage telomere resolvase TelK (25). It can also synapse replicated telomeres and catalyze the formation of Holliday junctions (29). The ability of ResT to promote hairpin fusion has been proposed as the mechanism underlying the ongoing genetic rearrangements that are a prominent feature of the B. burgdorferi genome (18, 28). Finally, B. burgdorferi ResT can tolerate a surprising amount of variation in its substrate (30, 31), a feature that is not shared by phage telomere resolvases (21). Although B. burgdorferi ResT appears to be more permissive with a greater scope of activities than other telomere resolvases, the sequences mediating most of its unique properties have not yet been identified.The B. burgdorferi genome contains a total of 19 distinct hairpin sequences, all of which must be resolved by ResT (31). These sequences can be classified into three groups based on the presence and positioning of the box 1 motif, which is a critical determinant of activity in phage and Borrelia telomere resolvases (see Fig. 1A) (21, 24, 30). A box 1-like motif is also found in many of the hairpin telomeres sequenced to date (6, 14, 3235), although its function in telomere resolution is unknown. The box 1 consensus sequence (TAT(a/t)AT) closely resembles the −10/Pribnow box and TATA box consensus sequences of prokaryotic and eukaryotic promoters (TATAAT and TATA(a/t)A(a/t), respectively), which undergo transient deformations that predispose them to melting (36) and are intrinsically bent and anisotropically flexible (37). Therefore, box 1 may facilitate nucleation of hairpin folding and/or may confer an intrinsic bend or flexibility to substrates that is important for the resolution reaction.Open in a separate windowFIGURE 1.Species-specific resolution of Type I and 2 telomeres. A, a schematic showing the three types of hairpin telomere found on the linear replicons of the B. burgdorferi genome (see Ref. 31). The box 1 sequence in Type 1 and 2 telomeres is situated 1 and 4 nucleotides away from the axis of symmetry, respectively, whereas Type 3 telomeres contain no clear box 1. B, a schematic illustrating the telomere resolution reaction substrate and products is shown along with two ethidium bromide-stained agarose gels showing telomere resolution assays. The gels show resolution kinetics for B. burgdorferi and B. hermsii ResT on Type 1 and 2 telomeres (plasmid substrates pYT1/lp17L and pYT92/chromL, respectively).B. burgdorferi ResT can resolve telomeres in which box 1 is located at positions 1 and 4 nucleotides away from the axis of symmetry (Type 1 and 2 telomeres, respectively), as well as AT-rich telomeres without a box 1 sequence (Type 3 telomeres) (see Fig. 1A) (30, 31). B. burgdorferi ResT cleaves telomeres at a fixed position relative to the axis of symmetry, independent of the location of box 1 (30). Positioning of the enzyme for cleavage in all telomere types is most likely driven by sequence-specific interactions between ResT domains 2 (catalytic) and/or 3 (C-terminal) and a fixed element upstream of box 1 that is positioned 14 nucleotides from the axis of symmetry in all Borrelia telomeres (box 3 and adjacent nucleotides) (see Figs. 1A and and2)2) (26, 30, 31). In contrast, box 1 and axis-flanking nucleotides are not involved in high affinity and/or sequence-specific interactions with ResT and require the ResT N-terminal domain for full protection in DNase footprinting assays (26, 27). The most likely candidate for interactions with box 1 and axis-flanking nucleotides is a Borrelia-specific hairpin-binding region in the N terminus, which is thought to promote a pre-hairpinning step involving strand opening at the axis (38).Open in a separate windowFIGURE 2.Alignment of 11 Borrelia ResT sequences. Shown is ClustalW2 alignment of ResT amino acid sequences from five Lyme disease Borrelia species (B. afzelii, B. spielmanii, B. valaisiana, B. garinii, and B. burgdorferi), five relapsing fever Borrelia species (B. turicatae, B. parkeri, B. hermsii, B. recurrentis, and B. duttonii), and one avian Borrelia species (B. anserina) (generated using ClustalW2 from the EBI web site) (19, 3942, 48, 49). The sequences for B. anserina, B. parkeri, and B. turicatae ResTs are reported for the first time in this study (respective GenBank accession numbers are FJ882620, FJ882621, and FJ882623). Sequences are arranged in order of similarity to neighboring sequences and are colored in JalView using the Zappo coloring scheme for identifying amino acids with similar physicochemical properties (50). Only residues that are identical in 100% of ResTs are indicated by colored shading. Arrows above the alignment indicate ResT domain boundaries identified by chymotrypsin digest, sequence comparison with other proteins, and HHsenser predictions (26, 51). The hairpin-binding motif found in cut-and-paste transposases is indicated beneath the alignment by white text on a black background (38). The positions corresponding to the active site residues in tyrosine recombinases, type IB topoisomerases, and TelK are indicated by blue asterisks below the sequence, with the active site tyrosine nucleophile at position 335 marked by a red asterisk (22, 25). The ringed black dot below position 326 indicates an amino acid in the active site region that differs in Lyme disease and relapsing fever ResTs. Sequences above the black line drawn between B. burgdorferi and B. turicatae are from Lyme disease Borrelia species; sequences below the black line are from relapsing fever Borrelia species. The ResT sequence from the avian Borrelia species B. anserina is shown at bottom.ResT from the relapsing fever Borrelia species Borrelia hermsii exhibits a more restricted substrate usage pattern in vitro when compared with ResT from the Lyme disease pathogen B. burgdorferi (39). Specifically, B. hermsii ResT is unable to efficiently resolve a Type 2 telomere. Therefore, B. burgdorferi ResT appears to be a more permissive enzyme than its relapsing fever counterpart. In this study, we investigated the basis for permissive substrate usage by B. burgdorferi ResT. Using a library of chimeric B. hermsii/B. burgdorferi ResTs, we mapped the sequence determinants in B. burgdorferi ResT that confer the ability to resolve multiple Type 2 telomeres. Surprisingly, this approach indicated that Type 2 telomere resolution was crucially regulated by a single proline residue located in a small Borrelia-specific insertion in the central catalytic region of ResT. The proline at this position was conserved in the ResTs from all Lyme disease Borrelia species but in none of the ResTs from relapsing fever Borrelia species, which were unable to efficiently resolve Type 2 telomeres in vitro. This study has identified a specific residue in ResT responsible for permissive substrate usage patterns.  相似文献   

20.
Protein–protein interactions (PPIs) are fundamental to the structure and function of protein complexes. Resolving the physical contacts between proteins as they occur in cells is critical to uncovering the molecular details underlying various cellular activities. To advance the study of PPIs in living cells, we have developed a new in vivo cross-linking mass spectrometry platform that couples a novel membrane-permeable, enrichable, and MS-cleavable cross-linker with multistage tandem mass spectrometry. This strategy permits the effective capture, enrichment, and identification of in vivo cross-linked products from mammalian cells and thus enables the determination of protein interaction interfaces. The utility of the developed method has been demonstrated by profiling PPIs in mammalian cells at the proteome scale and the targeted protein complex level. Our work represents a general approach for studying in vivo PPIs and provides a solid foundation for future studies toward the complete mapping of PPI networks in living systems.Protein–protein interactions (PPIs)1 play a key role in defining protein functions in biological systems. Aberrant PPIs can have drastic effects on biochemical activities essential to cell homeostasis, growth, and proliferation, and thereby lead to various human diseases (1). Consequently, PPI interfaces have been recognized as a new paradigm for drug development. Therefore, mapping PPIs and their interaction interfaces in living cells is critical not only for a comprehensive understanding of protein function and regulation, but also for describing the molecular mechanisms underlying human pathologies and identifying potential targets for better therapeutics.Several strategies exist for identifying and mapping PPIs, including yeast two-hybrid, protein microarray, and affinity purification mass spectrometry (AP-MS) (25). Thanks to new developments in sample preparation strategies, mass spectrometry technologies, and bioinformatics tools, AP-MS has become a powerful and preferred method for studying PPIs at the systems level (69). Unlike other approaches, AP-MS experiments allow the capture of protein interactions directly from their natural cellular environment, thus better retaining native protein structures and biologically relevant interactions. In addition, a broader scope of PPI networks can be obtained with greater sensitivity, accuracy, versatility, and speed. Despite the success of this very promising technique, AP-MS experiments can lead to the loss of weak/transient interactions and/or the reorganization of protein interactions during biochemical manipulation under native purification conditions. To circumvent these problems, in vivo chemical cross-linking has been successfully employed to stabilize protein interactions in native cells or tissues prior to cell lysis (1016). The resulting covalent bonds formed between interacting partners allow affinity purification under stringent and fully denaturing conditions, consequently reducing nonspecific background while preserving stable and weak/transient interactions (1216). Subsequent mass spectrometric analysis can reveal not only the identities of interacting proteins, but also cross-linked amino acid residues. The latter provides direct molecular evidence describing the physical contacts between and within proteins (17). This information can be used for computational modeling to establish structural topologies of proteins and protein complexes (1722), as well as for generating experimentally derived protein interaction network topology maps (23, 24). Thus, cross-linking mass spectrometry (XL-MS) strategies represent a powerful and emergent technology that possesses unparalleled capabilities for studying PPIs.Despite their great potential, current XL-MS studies that have aimed to identify cross-linked peptides have been mostly limited to in vitro cross-linking experiments, with few successfully identifying protein interaction interfaces in living cells (24, 25). This is largely because XL-MS studies remain challenging due to the inherent difficulty in the effective MS detection and accurate identification of cross-linked peptides, as well as in unambiguous assignment of cross-linked residues. In general, cross-linked products are heterogeneous and low in abundance relative to non-cross-linked products. In addition, their MS fragmentation is too complex to be interpreted using conventional database searching tools (17, 26). It is noted that almost all of the current in vivo PPI studies utilize formaldehyde cross-linking because of its membrane permeability and fast kinetics (1016). However, in comparison to the most commonly used amine reactive NHS ester cross-linkers, identification of formaldehyde cross-linked peptides is even more challenging because of its promiscuous nonspecific reactivity and extremely short spacer length (27). Therefore, further developments in reagents and methods are urgently needed to enable simple MS detection and effective identification of in vivo cross-linked products, and thus allow the mapping of authentic protein contact sites as established in cells, especially for protein complexes.Various efforts have been made to address the limitations of XL-MS studies, resulting in new developments in bioinformatics tools for improved data interpretation (2832) and new designs of cross-linking reagents for enhanced MS analysis of cross-linked peptides (24, 3339). Among these approaches, the development of new cross-linking reagents holds great promise for mapping PPIs on the systems level. One class of cross-linking reagents containing an enrichment handle have been shown to allow selective isolation of cross-linked products from complex mixtures, boosting their detectability by MS (3335, 4042). A second class of cross-linkers containing MS-cleavable bonds have proven to be effective in facilitating the unambiguous identification of cross-linked peptides (3639, 43, 44), as the resulting cross-linked products can be identified based on their characteristic and simplified fragmentation behavior during MS analysis. Therefore, an ideal cross-linking reagent would possess the combined features of both classes of cross-linkers. To advance the study of in vivo PPIs, we have developed a new XL-MS platform based on a novel membrane-permeable, enrichable, and MS-cleavable cross-linker, Azide-A-DSBSO (azide-tagged, acid-cleavable disuccinimidyl bis-sulfoxide), and multistage tandem mass spectrometry (MSn). This new XL-MS strategy has been successfully employed to map in vivo PPIs from mammalian cells at both the proteome scale and the targeted protein complex level.  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