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 26 毫秒
1.
Insulin receptor substrate-1 (IRS-1) is a highly phosphorylated adaptor protein critical to insulin and IGF-1 receptor signaling. Ser/Thr kinases impact the metabolic and mitogenic effects elicited by insulin and IGF-1 through feedback and feed forward regulation at the level of IRS-1. Ser/Thr residues of IRS-1 are also O-GlcNAc-modified, which may influence the phosphorylation status of the protein. To facilitate the understanding of the functional effects of O-GlcNAc modification on IRS-1-mediated signaling, we identified the sites of O-GlcNAc modification of rat and human IRS-1. Tandem mass spectrometric analysis of IRS-1, exogenously expressed in HEK293 cells, revealed that the C terminus, which is rich in docking sites for SH2 domain-containing proteins, was O-GlcNAc-modified at multiple residues. Rat IRS-1 was O-GlcNAc-modified at Ser914, Ser1009, Ser1036, and Ser1041. Human IRS-1 was O-GlcNAc-modified at Ser984 or Ser985, at Ser1011, and possibly at multiple sites within residues 1025–1045. O-GlcNAc modification at a conserved residue in rat (Ser1009) and human (Ser1011) IRS-1 is adjacent to a putative binding motif for the N-terminal SH2 domains of p85α and p85β regulatory subunits of phosphatidylinositol 3-kinase and the tyrosine phosphatase SHP2 (PTPN11). Immunoblot analysis using an antibody generated against human IRS-1 Ser1011 GlcNAc further confirmed the site of attachment and the identity of the +203.2-Da mass shift as β-N-acetylglucosamine. The accumulation of IRS-1 Ser1011 GlcNAc in HEPG2 liver cells and MC3T3-E1 preosteoblasts upon inhibition of O-GlcNAcase indicates that O-GlcNAcylation of endogenously expressed IRS-1 is a dynamic process that occurs at normal glucose concentrations (5 mm). O-GlcNAc modification did not occur at any known or newly identified Ser/Thr phosphorylation sites and in most cases occurred simultaneously with phosphorylation of nearby residues. These findings suggest that O-GlcNAc modification represents an additional layer of posttranslational regulation that may impact the specificity of effects elicited by insulin and IGF-1.Insulin receptor substrate-1 (IRS-1)1 is a highly phosphorylated adaptor protein critical to insulin and IGF-1 receptor signaling. Many of the metabolic and mitogenic effects elicited by insulin and IGF-1 are mediated and modulated by posttranslational modifications of IRS-1, and tight regulation at the posttranslational level is crucial for maintaining insulin sensitivity and controlling growth factor-induced proliferation. Following hormonal stimulation, IRS-1 is phosphorylated by the receptor tyrosine kinases creating SH2 domain docking sites for downstream binding partners including the p85 regulatory subunits of phosphatidylinositol 3-kinase, Grb2, and the tyrosine phosphatase SHP2 (PTPN11) (1). Binding of p85 phosphatidylinositol 3-kinase and Grb2 activate the PI3K/Akt and Ras-MAPK pathways, respectively, whereas binding of SHP2 results in tyrosine dephosphorylation and signal attenuation (2). Positive and negative feedback regulation by Ser/Thr kinases, such as Akt (3), c-Jun N-terminal kinase (JNK) (4), S6K (5), and ERK (6), impact the interactions of IRS-1 with SH2 domain proteins and the receptor thereby affecting the duration and outcome of the signal. IRS-1 has been described as a central node for the integration of information regarding the nutrient and stress status of the cell (7). This information is encoded by site-specific phosphorylation by a number of kinases that regulate the specificity of effects that are elicited following receptor stimulation. Many sites of Ser/Thr phosphorylation have been identified on IRS-1, and cross-talk among Tyr and Ser/Thr phosphorylations at specific residues is evidence of dynamic and complex posttranslational regulation (8, 9). Inappropriate phosphorylation of IRS-1 resulting in the disruption of interactions of IRS-1 with binding partners is implicated in the development of insulin resistance (10) and altered IGF-1 signaling in breast cancer tissue (11, 12).In addition to phosphorylation, Ser/Thr residues in IRS-1 are also dynamically modified by GlcNAc in a nutrient-responsive manner. As opposed to a negatively charged phosphate group, O-GlcNAcylation imparts a bulky, hydrophilic, electrostatically neutral moiety to Ser/Thr residues. The enzymes responsible for the incorporation and removal of the monosaccharide from proteins, O-GlcNAc-transferase and O-GlcNAcase, respectively, are localized in the cytoplasm and the nucleus of all eukaryotic cells (13, 14). Recent studies suggest that the activity of O-GlcNAc-transferase is regulated by insulin (15) and that localization of O-GlcNAc-transferase to the membrane is driven by direct association with phosphatidylinositide 3-phosphate (16). The abundance of O-GlcNAc modification on many proteins in the insulin signaling pathway increases with sustained high glucose and chronic insulin stimulation, and elevated O-GlcNAc modification of IRS-1 correlates with the development of insulin resistance in multiple cell types including 3T3-L1 adipocytes (17, 18), MIN6 pancreatic beta cells (19), Fao rat hepatoma cells (16), human aortic endothelial cells (20), and skeletal muscle (21). The impact of O-GlcNAcylation on insulin signaling and diabetic complications was reviewed recently (22, 23). The direct effect of O-GlcNAc modification on signaling via IRS-1 is not known because conditions that mimic those in the uncontrolled diabetic patient may also result in phosphorylation of IRS-1 at inhibitory sites (16, 24) and O-GlcNAc modification of other proteins in the insulin signaling pathway, such as the insulin receptor, Akt (18), FoxO (25), AMP-activated protein kinase (26), and β-catenin (17).To elucidate site-specific effects of O-GlcNAc modification on IRS-1-mediated signal transduction, we identified the sites of O-GlcNAc modification of rat and human IRS-1 by tandem mass spectrometry. To facilitate detection of the O-GlcNAc-modified peptides and assign the sites of modification, CID coupled with neutral loss-triggered MS3 and electron transfer dissociation (ETD) (27) tandem spectrometric approaches were used. Fragmentation of O-GlcNAc-modified peptides by ETD did not destroy the labile O-linkage (28) permitting direct detection of these peptides by the database searching algorithm ProteinProspector2 (29). O-GlcNAc modification occurred in close proximity to multiple SH2 domain binding motifs and within a region of IRS-1 shown previously to interact with the insulin and IGF-1 receptors (30).  相似文献   

2.
3.
4.
5.
We report an integrated pipeline for efficient serum glycoprotein biomarker candidate discovery and qualification that may be used to facilitate cancer diagnosis and management. The discovery phase used semi-automated lectin magnetic bead array (LeMBA)-coupled tandem mass spectrometry with a dedicated data-housing and analysis pipeline; GlycoSelector (http://glycoselector.di.uq.edu.au). The qualification phase used lectin magnetic bead array-multiple reaction monitoring-mass spectrometry incorporating an interactive web-interface, Shiny mixOmics (http://mixomics-projects.di.uq.edu.au/Shiny), for univariate and multivariate statistical analysis. Relative quantitation was performed by referencing to a spiked-in glycoprotein, chicken ovalbumin. We applied this workflow to identify diagnostic biomarkers for esophageal adenocarcinoma (EAC), a life threatening malignancy with poor prognosis in the advanced setting. EAC develops from metaplastic condition Barrett''s esophagus (BE). Currently diagnosis and monitoring of at-risk patients is through endoscopy and biopsy, which is expensive and requires hospital admission. Hence there is a clinical need for a noninvasive diagnostic biomarker of EAC. In total 89 patient samples from healthy controls, and patients with BE or EAC were screened in discovery and qualification stages. Of the 246 glycoforms measured in the qualification stage, 40 glycoforms (as measured by lectin affinity) qualified as candidate serum markers. The top candidate for distinguishing healthy from BE patients'' group was Narcissus pseudonarcissus lectin (NPL)-reactive Apolipoprotein B-100 (p value = 0.0231; AUROC = 0.71); BE versus EAC, Aleuria aurantia lectin (AAL)-reactive complement component C9 (p value = 0.0001; AUROC = 0.85); healthy versus EAC, Erythroagglutinin Phaseolus vulgaris (EPHA)-reactive gelsolin (p value = 0.0014; AUROC = 0.80). A panel of 8 glycoforms showed an improved AUROC of 0.94 to discriminate EAC from BE. Two biomarker candidates were independently verified by lectin magnetic bead array-immunoblotting, confirming the validity of the relative quantitation approach. Thus, we have identified candidate biomarkers, which, following large-scale clinical evaluation, can be developed into diagnostic blood tests. A key feature of the pipeline is the potential for rapid translation of the candidate biomarkers to lectin-immunoassays.Biomarkers play a central role in health care by enabling accurate diagnosis and prognosis; hence there is extensive research on the identification and development of novel biomarkers. However, despite numerous biomarker publications over the years (1), only a handful of new cancer biomarkers have successfully completed the journey from discovery, qualification, to verification and validation (24). One possible way to overcome this challenge is to develop an integrated biomarker pipeline that facilitates the smooth and successful transition from discovery to validation (510). The first and foremost consideration in an integrated pipeline is the sample source. In general, most of the proteomics based workflows use tissues or proximal fluids during the discovery phase, with the goal of extending the findings to plasma. Although this approach avoid the high complexity serum/plasma proteome and the associated requisite multi-dimensional sample separation in discovery stages, it often leads to failure when the candidates are not detected in plasma because of the limited sensitivity of the available analytical methods, or the absence of candidates in the plasma (11). To overcome this pitfall, we have developed an integrated glycoprotein biomarker pipeline, which can simply and rapidly isolate glycosylated proteins from serum to enable high throughput analysis of differentially glycosylated proteins in discovery and qualification stages.The workflow utilizes naturally occurring glycan binding proteins, lectins, in a semi-automated high throughput workflow called lectin magnetic bead array-tandem mass spectrometry (LeMBA-MS/MS)1 (12, 13). Although lectins have been well-utilized in glycobiology and biomarker discovery (1417), the LeMBA-MS/MS workflow demonstrates several unique features. First, serum glycoproteins are isolated in a single-step using 20 individual lectin-coated magnetic beads in microplate format. Second, we have optimized the concentrations of salts and detergents for sample denaturation to avoid co-isolation of protein complexes without adversely affecting lectin pull-down efficiency. Third, a liquid handler is used for sample processing to facilitate high-throughput screening and increase reproducibility. In addition, we have optimized on-bead trypsin digestion and incorporated lectin-exclusion lists during nano-LC-MS/MS to identify nonglycosylated peptides from the isolated glycoproteins. With these innovations, LeMBA-MS/MS demonstrates nanomolar sensitivity and linearity, and applicability across species (12). Compared with existing single, serial or multi-lectin affinity chromatography (18, 19), LeMBA-MS/MS offers the capability to simultaneously screen 20 lectins in a semi-automated, high throughput manner. On the other hand, because LeMBA-MS/MS identifies the nonglycosylated peptides, it cannot be used for glycan site assignment and glycan structure elucidation (2023). However, the main advantage of LeMBA, we believe, is as a part of an integrated translational biomarker pipeline leading to lectin immunoassays. The lack of glycan structure details is not critical for clinical translation, as exemplified by the alpha-fetoprotein-L3 (AFP-L3) test, which measures the Lens culinaris agglutinin (LCA) binding fraction of serum alpha-fetoprotein (24, 25), and has been approved by the U.S. Food and Drug Administration for detection of hepatocellular carcinoma.In this study, we report the extension of the glycoprotein biomarker pipeline to the qualification phase with LeMBA-MRM-MS, and introduce statistical analysis pipelines GlycoSelector (http://glycoselector.di.uq.edu.au/) and Shiny mixOmics (http://mixomics-projects.di.uq.edu.au/Shiny) for the discovery and qualification phases, respectively. The utility of this integrated serum glycoprotein biomarker pipeline is demonstrated using esophageal adenocarcinoma (EAC) with unmet clinical need for an in vitro diagnostic test. EAC is a lethal malignancy of the lower esophagus with very poor 5-year survival rate of less than 25% (26). EAC is becoming increasingly common and its incidence is associated with the prevalent precursor metaplastic condition Barrett''s esophagus (BE), but with a low annual conversion rate of up to 1% (27). A common set of risk factors are described for BE and EAC, include gastroesophageal reflux disease (GERD), obesity, male gender, and smoking (28, 29). The current endoscopy-biopsy based diagnosis is invasive and costly, leading to an ineffective surveillance program. A blood test employing serum biomarkers that can distinguish patients with EAC from those with either BE or healthy tissue would, potentially, change the paradigm for the way in which BE and EAC are managed in the population (30). Serum glycan profiling studies have shown differential expression of glycan structures between healthy, BE, early dysplastic and EAC patients (3135). However, diagnostic serum glycoproteins showing differential glycosylation hence differential lectin binding remain to be discovered, making it a suitable disease model for this study.  相似文献   

6.
7.
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   

8.
9.
Cross-linking/mass spectrometry resolves protein–protein interactions or protein folds by help of distance constraints. Cross-linkers with specific properties such as isotope-labeled or collision-induced dissociation (CID)-cleavable cross-linkers are in frequent use to simplify the identification of cross-linked peptides. Here, we analyzed the mass spectrometric behavior of 910 unique cross-linked peptides in high-resolution MS1 and MS2 from published data and validate the observation by a ninefold larger set from currently unpublished data to explore if detailed understanding of their fragmentation behavior would allow computational delivery of information that otherwise would be obtained via isotope labels or CID cleavage of cross-linkers. Isotope-labeled cross-linkers reveal cross-linked and linear fragments in fragmentation spectra. We show that fragment mass and charge alone provide this information, alleviating the need for isotope-labeling for this purpose. Isotope-labeled cross-linkers also indicate cross-linker-containing, albeit not specifically cross-linked, peptides in MS1. We observed that acquisition can be guided to better than twofold enrich cross-linked peptides with minimal losses based on peptide mass and charge alone. By help of CID-cleavable cross-linkers, individual spectra with only linear fragments can be recorded for each peptide in a cross-link. We show that cross-linked fragments of ordinary cross-linked peptides can be linearized computationally and that a simplified subspectrum can be extracted that is enriched in information on one of the two linked peptides. This allows identifying candidates for this peptide in a simplified database search as we propose in a search strategy here. We conclude that the specific behavior of cross-linked peptides in mass spectrometers can be exploited to relax the requirements on cross-linkers.Cross-linking/mass spectrometry extends the use of mass-spectrometry-based proteomics from identification (1, 2), quantification (3), and characterization of protein complexes (4) into resolving protein structures and protein–protein interactions (58). Chemical reagents (cross-linkers) covalently connect amino acid pairs that are within a cross-linker-specific distance range in the native three-dimensional structure of a protein or protein complex. A cross-linking/mass spectrometry experiment is typically conducted in four steps: (1) cross-linking of the target protein or complex, (2) protein digestion (usually with trypsin), (3) LC-MS analysis, and (4) database search. The digested peptide mixture consists of linear and cross-linked peptides, and the latter can be enriched by strong cation exchange (9) or size exclusion chromatography (10). Cross-linked peptides are of high value as they provide direct information on the structure and interactions of proteins.Cross-linked peptides fragment under collision-induced dissociation (CID) conditions primarily into b- and y-ions, as do their linear counterparts. An important difference regarding database searches between linear and cross-linked peptides stems from not knowing which peptides might be cross-linked. Therefore, one has to consider each single peptide and all pairwise combinations of peptides in the database. Having n peptides leads to (n2 + n)/2 possible pairwise combinations. This leads to two major challenges: With increasing size of the database, search time and the risk of identifying false positives increases. One way of circumventing these problems is to use MS2-cleavable cross-linkers (11, 12), at the cost of limited experimental design and choice of cross-linker.In a first database search approach (13), all pairwise combinations of peptides in a database were considered in a concatenated and linearized form. Thereby, all possible single bond fragments are considered in one of the two database entries per peptide pair, and the cross-link can be identified by a normal protein identification algorithm. Already, the second search approach split the peptides for the purpose of their identification (14). Linear fragments were used to retrieve candidate peptides from the database that are then matched based on the known mass of the cross-linked pair and scored as a pair against the spectrum. Isotope-labeled cross-linkers were used to sort the linear and cross-linked fragments apart. Many other search tools and approaches have been developed since (10, 1519); see (20) for a more detailed list, at least some of which follow the general idea of an open modification search (2124).As a general concept for open modification search of cross-linked peptides, cross-linked peptides represent two peptides, each with an unknown modification given by the mass of the other peptide and the cross-linker. One identifies both peptides individually and then matches them based on knowing the mass of cross-linked pair (14, 22, 24). Alternatively, one peptide is identified first and, using that peptide and the cross-linker as a modification mass, the second peptide is identified from the database (21, 23). An important element of the open modification search approach is that it essentially converts the quadratic search space of the cross-linked peptides into a linear search space of modified peptides. Still, many peptides and many modification positions have to be considered, especially when working with large databases or when using highly reactive cross-linkers with limited amino acid selectivity (25).We hypothesize that detailed knowledge of the fragmentation behavior of cross-linked peptides might reveal ways to improve the identification of cross-linked peptides. Detailed analyses of the fragmentation behavior of linear peptides exist (2628), and the analysis of the fragmentation behavior of cross-linked peptides has guided the design of scores (24, 29). Further, cross-link-specific ions have been observed from higher energy collision dissociation (HCD) data (30). Isotope-labeled cross-linkers are used to distinguish cross-linked from linear fragments, generally in low-resolution MS2 of cross-linked peptides (14).We compared the mass spectrometric behavior of cross-linked peptides to that of linear peptides, using 910 high-resolution fragment spectra matched to unique cross-linked peptides from multiple different public datasets at 5% peptide-spectrum match (PSM)1 false discovery rate (FDR). In addition, we repeated all experiments with a larger sample set that contains 8,301 spectra—also including data from ongoing studies from our lab (Supplemental material S9-S12). This paper presents the mass spectrometric signature of cross-linked peptides that we identified in our analysis and the resulting heuristics that are incorporated into an integrated strategy for the analysis and identification of cross-linked peptides. We present computational strategies that indicate the possibility of alleviating the need for mass-spectrometrically restricted cross-linker choice.  相似文献   

10.
11.
Site-specific glycosylation analysis is key to investigate structure-function relationships of glycoproteins, e.g. in the context of antigenicity and disease progression. The analysis, though, is quite challenging and time consuming, in particular for O-glycosylated proteins. In consequence, despite their clinical and biopharmaceutical importance, many human blood plasma glycoproteins have not been characterized comprehensively with respect to their O-glycosylation. Here, we report on the site-specific O-glycosylation analysis of human blood plasma glycoproteins. To this end pooled human blood plasma of healthy donors was proteolytically digested using a broad-specific enzyme (Proteinase K), followed by a precipitation step, as well as a glycopeptide enrichment and fractionation step via hydrophilic interaction liquid chromatography, the latter being optimized for intact O-glycopeptides carrying short mucin-type core-1 and -2 O-glycans, which represent the vast majority of O-glycans on human blood plasma proteins. Enriched O-glycopeptide fractions were subjected to mass spectrometric analysis using reversed-phase liquid chromatography coupled online to an ion trap mass spectrometer operated in positive-ion mode. Peptide identity and glycan composition were derived from low-energy collision-induced dissociation fragment spectra acquired in multistage mode. To pinpoint the O-glycosylation sites glycopeptides were fragmented using electron transfer dissociation. Spectra were annotated by database searches as well as manually. Overall, 31 O-glycosylation sites and regions belonging to 22 proteins were identified, the majority being acute-phase proteins. Strikingly, also 11 novel O-glycosylation sites and regions were identified. In total 23 O-glycosylation sites could be pinpointed. Interestingly, the use of Proteinase K proved to be particularly beneficial in this context. The identified O-glycan compositions most probably correspond to mono- and disialylated core-1 mucin-type O-glycans (T-antigen). The developed workflow allows the identification and characterization of the major population of the human blood plasma O-glycoproteome and our results provide new insights, which can help to unravel structure-function relationships. The data were deposited to ProteomeXchange PXD003270.Human blood plasma harbors arguably the most complex yet also the most informative proteome present in the human body (1). A significant impact on its clinical relevance and diagnostic potential is attributed to the features and functions of a plethora of proteins (60–80 mg protein per ml plasma), covering a dynamic concentration range of more than ten orders of magnitude (2). The majority, that is 99%, of these proteins are classical blood plasma proteins, like albumins, (immuno)globulins, clotting factors, and proteins of the complement system; however, also a lower abundant but—no less meaningful—fraction of nonclassical proteins is present that comprises a multitude of cytokines as well as tissue leakage proteins. Several clinical studies could show that qualitative and quantitative alterations of these proteins (and peptides)—analyzed individually or in their entirety as a proteome (or peptidome)—can directly reflect pathophysiological states, and can serve as biomarkers for the onset and progression of a number of diseases (35). In recent years the focus of in-depth analyses of the human blood plasma proteome has evolved from the identification and quantification of the entire proteome (or peptidome) (610) toward the analysis of subproteomes like the interactome (11), phosphoproteome (12, 13) or the glycoproteome (14). The latter has received particular interest in recent years, because the majority of blood plasma proteins is N- and/or O-glycosylated (2). Although the comprehensive analysis of the N-glycoproteome is already quite advanced (15), even in complex samples like human blood plasma (16, 17), similar analyses of the O-glycoproteome - though arguably equally important and relevant - are still lagging behind. The most ubiquitously found and functionally relevant form of O-glycosylation, as shown by a number of O-glycan-related (clinical) studies (1823), is the mucin-type O-glycosyation (O-GalNAc), in particular the core-1 and core-2 types (24, 25). The predominantly clustered occurrence of mucin-type O-glycans on proteins is described to confer overall stability and proteolytic protection (26). Apart from this global impact, recent studies could link the presence of O-glycans in the proximity of regulatory domains to proteolysis events involved in protein maturation (proprotein-convertase-processing) (27). To better understand these protective and regulatory capabilities and to move the mucin-type O-glycoproteome from form to function comprehensive site-specific O-glycosylation analyses are required.One of the main obstacles in site-specific mucin-type O-glycosylation analyses relates to the lack of a predictable O-glycan consensus-motif within the peptide backbone as it can be found for N-glycans (28). The initial attachment of the N-acetylgalactosamine monosaccharide to the hydroxyl group of either serine or threonine, but also to tyrosine or hydroxylysine, is governed by a family of 20 distinct polypeptide GalNAc-transferase isoenzymes (GalNAc-Ts) with different but partially overlapping peptide specificities and tissue expression patterns. This dynamic regulation, in turn, contributes to the complexity of the mucin-type O-glycoproteome. However, previous studies could show that mucin-type O-glycans are primarily attached to serine or threonine in regions with a high content of serine, threonine and proline (Ser/Thr-X-X-Pro, Ser/Thr-P and Pro-Ser/Thr) (29, 30). As O-glycosylation is a postfolding event, taking place in the Golgi apparatus, the attachment is depended on protein surface accessibility and is thus predominantly found in coil, turn, and linker regions (31). Additional confounding factors during mucin-type O-glycosylation analyses are the clustered occurrence of O-glycans and the lack of a universal endo-O-glycosidase that enables the release of intact O-glycans from the proteins; though, chemical O-glycan release methods do exist (28).Mass spectrometry has proven to be the core technique in site-specific N- and O-glycosylation analyses. A generic O-glycoproteomic workflow usually starts with the isolation, enrichment or prefractionation of a single glycoprotein or a group of glycoproteins. In subsequent steps, (glyco)peptides are generated by proteolytic digestion primarily using specific proteases like trypsin. Apart from this, also broad- and nonspecific proteases like Proteinase K or Pronase E were successfully employed in recent years (3234). Essential to nearly every glycoproteomic approach is the removal of high-abundant and interfering nonglycosylated peptides by selective enrichment of the usually lower abundant glycopeptides. The repertoire of glycopeptide enrichment and separation techniques covers different solid phase extraction and chromatography based methods such as hydrophilic liquid interaction chromatography (HILIC) (35, 36). The most frequently used setup for the measurement of enriched (glyco)peptides is liquid chromatography (LC)1 coupled online to electrospray ionization tandem mass spectrometry (LC-ESI-MS/MS). Recent advances in instrumentation, in particular the development of electron-transfer/electron-capture dissociation (ETD/ECD) (37, 38), and high resolution orbital mass analyzers, have paved the way for the mapping of thousands of occupied N- and O-glycosylation sites as recently shown (17, 27). Combined workflows using ETD/ECD fragmentation along with (multistage, MSn) fragmentation with high- and/or low collisional induced dissociation energy (HCD/CID) can provide compositional (structural) information on the glycan moiety as well as information on the peptide sequence and the glycosylation site (39, 40). Recent advances in mass spectrometry driven O-glycoproteomics have been reviewed in detail elsewhere (41, 42). Owing to the amount and complexity of O-glycoproteomic data a number of bioinformatic software tools for the prediction of mucin-type O-glycosylation sites (27) as well as for the database assisted interpretation and annotation of glycan and glycopeptide fragment spectra have been developed (43, 44). Moreover, reporting guidelines for collecting, sharing, integrating, and interpreting mass spectrometry based glycomics data have been specified by the MIRAGE consortium (minimum information required for a glycomics experiment) (45, 46).The aim of our study was to develop a glycoproteomic workflow that allows the explorative nontargeted analysis of O-glycosylated human blood plasma proteins, which are known to carry mainly short mono- and disialylated mucin-type core-1 and -2 O-glycans. To achieve this, we have combined O-glycopeptide selective offline-HILIC fractionation of Proteinase K digested peptides with nano-reversed-phase liquid chromatography coupled online to multistage ion-trap mass spectrometry (nanoRP-LC-ESI-IT-MS: CID-MS2/-MS3, ETD-MS2). The workflow has been applied to investigate the mucin-type O-glycoproteome of a pooled blood plasma sample derived from 20 healthy donors. Based on the mass spectrometric analysis of intact O-glycopeptides, we were able to characterize the O-glycosylation (i.e. peptide, site, and attached O-glycans) of a number of major human blood glycoproteins, including many acute phase proteins such as fibrinogen and plasminogen. Overall, the site-specific glycosylation analysis of human blood plasma glycopeptides revealed exclusively mono- and disialylated core-1 mucin-type O-glycopeptides. Interestingly, also a few novel O-glycosylation sites could be identified.  相似文献   

12.
13.
Glycoprotein structure determination and quantification by MS requires efficient isolation of glycopeptides from a proteolytic digest of complex protein mixtures. Here we describe that the use of acids as ion-pairing reagents in normal-phase chromatography (IP-NPLC) considerably increases the hydrophobicity differences between non-glycopeptides and glycopeptides, thereby resulting in the reproducible isolation of N-linked high mannose type and sialylated glycopeptides from the tryptic digest of a ribonuclease B and fetuin mixture. The elution order of non-glycopeptides relative to glycopeptides in IP-NPLC is predictable by their hydrophobicity values calculated using the Wimley-White water/octanol hydrophobicity scale. O-linked glycopeptides can be efficiently isolated from fetuin tryptic digests using IP-NPLC when N-glycans are first removed with PNGase. IP-NPLC recovers close to 100% of bacterial N-linked glycopeptides modified with non-sialylated heptasaccharides from tryptic digests of periplasmic protein extracts from Campylobacter jejuni 11168 and its pglD mutant. Label-free nano-flow reversed-phase LC-MS is used for quantification of differentially expressed glycopeptides from the C. jejuni wild-type and pglD mutant followed by identification of these glycoproteins using multiple stage tandem MS. This method further confirms the acetyltransferase activity of PglD and demonstrates for the first time that heptasaccharides containing monoacetylated bacillosamine are transferred to proteins in both the wild-type and mutant strains. We believe that IP-NPLC will be a useful tool for quantitative glycoproteomics.Protein glycosylation is a biologically significant and complex post-translational modification, involved in cell-cell and receptor-ligand interactions (14). In fact, clinical biomarkers and therapeutic targets are often glycoproteins (59). Comprehensive glycoprotein characterization, involving glycosylation site identification, glycan structure determination, site occupancy, and glycan isoform distribution, is a technical challenge particularly for quantitative profiling of complex protein mixtures (1013). Both N- and O-glycans are structurally heterogeneous (i.e. a single site may have different glycans attached or be only partially occupied). Therefore, the MS1 signals from glycopeptides originating from a glycoprotein are often weaker than from non-glycopeptides. In addition, the ionization efficiency of glycopeptides is low compared with that of non-glycopeptides and is often suppressed in the presence of non-glycopeptides (1113). When the MS signals of glycopeptides are relatively high in simple protein digests then diagnostic sugar oxonium ion fragments produced by, for example, front-end collisional activation can be used to detect them. However, when peptides and glycopeptides co-elute, parent ion scanning is required to selectively detect the glycopeptides (14). This can be problematic in terms of sensitivity, especially for detecting glycopeptides in digests of complex protein extracts.Isolation of glycopeptides from proteolytic digests of complex protein mixtures can greatly enhance the MS signals of glycopeptides using reversed-phase LC-ESI-MS (RPLC-ESI-MS) or MALDI-MS (1524). Hydrazide chemistry is used to isolate, identify, and quantify N-linked glycopeptides effectively, but this method involves lengthy chemical procedures and does not preserve the glycan moieties thereby losing valuable information on glycan structure and site occupancy (1517). Capturing glycopeptides with lectins has been widely used, but restricted specificities and unspecific binding are major drawbacks of this method (1821). Under reversed-phase LC conditions, glycopeptides from tryptic digests of gel-separated glycoproteins have been enriched using graphite powder medium (22). In this case, however, a second digestion with proteinase K is required for trimming down the peptide moieties of tryptic glycopeptides so that the glycopeptides (typically <5 amino acid residues) essentially resemble the glycans with respect to hydrophilicity for subsequent separation. Moreover, the short peptide sequences of the proteinase K digest are often inadequate for de novo sequencing of the glycopeptides.Glycopeptide enrichment under normal-phase LC (NPLC) conditions has been demonstrated using various hydrophilic media and different capture and elution conditions (2328). NPLC allows either direct enrichment of peptides modified by various N-linked glycan structures using a ZIC®-HILIC column (2327) or targeting sialylated glycopeptides using a titanium dioxide micro-column (28). However, NPLC is neither effective for enriching less hydrophilic glycopeptides, e.g. the five high mannose type glycopeptides modified by 7–11 monosaccharide units from a tryptic digest of ribonuclease b (RNase B), nor for enriching O-linked glycopeptides of bovine fetuin using a ZIC-HILIC column (23). The use of Sepharose medium for enriching glycopeptides yielded only modest recovery of glycopeptides (28). In addition, binding of hydrophilic non-glycopeptides with these hydrophilic media contaminates the enriched glycopeptides (23, 28).We have recently developed an ion-pairing normal-phase LC (IP-NPLC) method to enrich glycopeptides from complex tryptic digests using Sepharose medium and salts or bases as ion-pairing reagents (29). Though reasonably effective the technique still left room for significant improvement. For example, the method demonstrated relatively modest glycopeptide selectivity, providing only 16% recovery for high mannose type glycopeptides (29). Here we report on a new IP-NPLC method using acids as ion-pairing reagents and polyhydroxyethyl aspartamide (A) as the stationary phase for the effective isolation of tryptic glycopeptides. The method was developed and evaluated using a tryptic digest of RNase B and fetuin mixture. In addition, we demonstrate that O-linked glycopeptides can be effectively isolated from a fetuin tryptic digest by IP-NPLC after removal of the N-linked glycans by PNGase F.The new IP-NPLC method was used to enrich N-linked glycopeptides from the tryptic digests of protein extracts of wild-type (wt) and PglD mutant strains of Campylobacter jejuni NCTC 11168. C. jejuni has a unique N-glycosylation system that glycosylates periplasmic and inner membrane proteins containing the extended N-linked sequon, D/E-X-N-X-S/T, where X is any amino acid other than proline (3032). The N-linked glycan of C. jejuni has been previously determined to be GalNAc-α1,4-GalNAc-α1,4-[Glcβ1,3]-GalNAc-α1,4-GalNAc-α1,4-GalNAc-α1,3-Bac-β1 (BacGalNAc5Glc residue mass: 1406 Da), where Bac is 2,4-diacetamido-2,4,6-trideoxyglucopyranose (30). In addition, the glycan structure of C. jejuni is conserved, unlike in eukaryotic systems (3032). IP-NPLC recovered close to 100% of the bacterial N-linked glycopeptides with virtually no contamination of non-glycopeptides. Furthermore, we demonstrate for the first time that acetylation of bacillosamine is incomplete in the wt using IP-NPLC and label-free MS.  相似文献   

14.
15.
The lack of consensus sequence, common core structure, and universal endoglycosidase for the release of O-linked oligosaccharides makes O-glycosylation more difficult to tackle than N-glycosylation. Structural elucidation by mass spectrometry is usually inconclusive as the CID spectra of most glycopeptides are dominated by carbohydrate-related fragments, preventing peptide identification. In addition, O-linked structures also undergo a gas-phase rearrangement reaction, which eliminates the sugar without leaving a telltale sign at its former attachment site. In the present study we report the enrichment and mass spectrometric analysis of proteins from bovine serum bearing Galβ1–3GalNAcα (mucin core-1 type) structures and the analysis of O-linked glycopeptides utilizing electron transfer dissociation and high resolution, high mass accuracy precursor ion measurements. Electron transfer dissociation (ETD) analysis of intact glycopeptides provided sufficient information for the identification of several glycosylation sites. However, glycopeptides frequently feature precursor ions of low charge density (m/z > ∼850) that will not undergo efficient ETD fragmentation. Exoglycosidase digestion was utilized to reduce the mass of the molecules while retaining their charge. ETD analysis of species modified by a single GalNAc at each site was significantly more successful in the characterization of multiply modified molecules. We report the unambiguous identification of 21 novel glycosylation sites. We also detail the limitations of the enrichment method as well as the ETD analysis.Glycosylation is among the most prevalent post-translational modifications of proteins; it is estimated that over half of all proteins undergo glycosylation during their lifespan (1). Glycosylation of secreted proteins and the extracellular part of membrane proteins occurs in the endoplasmic reticulum and the contiguous Golgi complex. The side chains of Trp, Asn, and Thr/Ser residues can be modified, termed as C-, N-, and O-glycosylation, respectively (2, 3). In addition, O-glycosylation also occurs within the nucleus and the cytosol: a single GlcNAc residue modifies Ser and Thr residues. O-GlcNAc glycosylation fulfills a regulatory/signaling function similar to phosphorylation (4).From an analytical point of view, C-glycosylation is the simplest. A consensus sequence has been defined: WXXW where the first Trp is modified, and the modification, a Man moiety, readily survives sample preparation and mass spectrometric analysis, including collisional activation (5). Investigation of N-glycosylation is also facilitated by several factors. First, N-glycosylation again has a well defined consensus sequence: NX(S/T/C) where the middle amino acid cannot be Pro (6). Second, there is a universal core glycan structure: GlcNAc2Man3; and this core is conserved across species. Third, a specific endoglycosidase, peptide N-glycosidase F, has been identified. This enzyme cleaves the carbohydrate structure from the peptide, leaving behind a diagnostic sign: the Asn residue is hydrolyzed to Asp, inducing a mass shift of +1 Da. By contrast, analysis of O-glycosylation is hampered by a lack of (i) a consensus sequence, (ii) a universal core structure, and (iii) a universal endoglycosidase or gentle chemical hydrolysis method to facilitate analysis.Glycosylation shows a high degree of species and tissue specificity; the same site may be modified by a wide variety of different glycan structures, and unmodified variants of the protein may occur simultaneously (79). Disease and physiological changes also may alter the glycosylation pattern (1012). The biological role(s) of glycosylation has been studied extensively (1315), although such studies are seriously hampered by the difficulties of glycosylation analysis.Most secreted proteins are glycosylated; and thus, mammalian serum is rich in glycoproteins. On the other hand, O-linked glycoproteins represent a small percentage of the serum protein content. Glycoproteins may display a befuddling heterogeneity both in site specificity and site occupancy. Thus, the enrichment of modified proteins or peptides is necessary for their characterization, and different techniques have been tested for this purpose. Lectin affinity chromatography is a popular method for selective isolation of glycoproteins and glycopeptides. Concanavalin A can be used to isolate oligomannose type glycopeptides (16), wheat germ agglutinin is applied for GlcNAc-containing compounds (16, 17), and jacalin is selective for core-1 type O-glycopeptides (18, 19). Lectins with preferential affinity for fucosylated and sialylated structures can also be utilized (12). Non-selective capture of glycopeptides can be performed using hydrophilic interaction chromatography (20, 21) or size exclusion chromatography (22). A recent approach applies porous graphite columns for semiselective enrichment (23), whereas the acidic character of sialylated glycopeptides has also been exploited via titanium dioxide-mediated enrichment (24). Finally vicinal cis-diols can be selectively captured using boronic acid derivatives (2527). All methods described here provide some glycopeptide enrichment from non-glycosylated peptide background, but all also suffer from significant non-selective binding. N-Linked glycoproteins may also be selectively captured on hydrazide resin following periodate oxidation (28). This approach requires enzymatic deglycosylation to release the captured peptides for analysis, therefore excluding the determination of the carbohydrate structure.Intact glycopeptide characterization still represents a significant challenge. Edman degradation, either alone or in combination with mass spectrometry, has been utilized for such tasks (29, 30). CID analysis of O-linked glycopeptides has limited utility. (i) MS/MS analysis cannot differentiate between the isomeric carbohydrate units and usually does not reveal the linkage positions and the configuration of the glycosidic bonds. (ii) Such spectra are typically dominated by abundant product ions associated with carbohydrate fragmentation, namely non-reducing end oxonium ions and product ions formed via sequential neutral losses of sugar residues from the precursor ions. (iii) The glycan is cleaved from the peptide via a gas-phase rearrangement reaction, and as a result the peptide itself and most peptide fragments (if any) are detected partially or completely deglycosylated (3133). Recently a different approach, the combination of positive and negative ion mode infrared multiphoton dissociation, was found to provide conclusive structural assignment for some O-linked glycopeptides (34). However, two novel MS/MS techniques, electron capture dissociation (ECD),1 which is performed in FT-ICR mass spectrometers (35), and electron transfer dissociation (ETD), which is performed in various ion trapping devices (36), may represent the real breakthrough. In both cases an electron is transferred to multiply protonated peptide cations, triggering peptide fragmentation at the covalent bond between the amino group and the α-carbon, producing mostly c and radical z· product ions while leaving the side chains intact. ETD is typically more efficient than ECD and thus leads to more comprehensive fragmentation. In addition, ETD can be performed in ion traps and thus, at a higher sensitivity level, especially in a linear ion trap. Because it has been observed that there are instances when the electron transfer is efficient and still no significant fragmentation occurs, ETD is usually combined with supplementary (and gentle) CID activation (37). O-Glycosylation analysis using these new dissociative techniques has been investigated (38, 39). However, because of the complexity of extracellular O-glycosylation, analysis of complex mixtures is rarely attempted (18), and the above techniques are usually restricted to the analysis of purified proteins.In this study we present the analysis of secreted O-linked glycopeptides. Lectin (jacalin) affinity chromatography was used to achieve some enrichment of core-1 O-GalNAcα type carbohydrate-carrying glycopeptides from bovine serum. The glycopeptide fractions were subjected to CID and ETD analysis. These experiments were performed on a linear ion trap-Orbitrap hybrid mass spectrometer (40). The Orbitrap delivered high resolution, high mass accuracy for the precursor ions, whereas the linear trap provided high sensitivity MS/MS analyses. Some fractions were also subjected to sequential exoglycosidase digestions, and glycopeptides retaining only the proximal GalNAc residues were analyzed. ProteinProspector v5.2.1, developed to accommodate ETD product ion spectra, aided data interpretation (41). We identified 26 glycosylation sites from bovine serum unambiguously; 21 of these sites have never been reported by any other study. No other single study to date has yielded so much information about O-linked glycosylation sites.  相似文献   

16.
Human serum albumin is one of the most abundant plasma proteins that readily undergoes glycation, thus glycated albumin has been suggested as an additional marker for monitoring glycemic status. Hitherto, only Amadori-modified peptides of albumin were quantified. In this study, we report the construction of fragment ion library for Amadori-modified lysine (AML), N(ε)-(carboxymethyl)lysine (CML)-, and N(ε)-(carboxyethyl)lysine (CEL)-modified peptides of the corresponding synthetically modified albumin using high resolution accurate mass spectrometry (HR/AM). The glycated peptides were manually inspected and validated for their modification. Further, the fragment ion library was used for quantification of glycated peptides of albumin in the context of diabetes. Targeted Sequential Window Acquisition of all THeoretical Mass Spectra (SWATH) analysis in pooled plasma samples of control, prediabetes, diabetes, and microalbuminuria, has led to identification and quantification of 13 glycated peptides comprised of four AML, seven CML, and two CEL modifications, representing nine lysine sites of albumin. Five lysine sites namely K549, K438, K490, K88, and K375, were observed to be highly sensitive for glycation modification as their respective m/z showed maximum fold change and had both AML and CML modifications. Thus, peptides involving these lysine sites could be potential novel markers to assess the degree of glycation in diabetes.Diabetes is a complex metabolic disorder characterized by prolonged hyperglycemia resulting from defects in insulin secretion, insulin action, or both, leading to abnormalities in carbohydrate, fat, and protein metabolism (1). According to the projection by the International Diabetes Foundation, around 592 million people will be affected by diabetes by the year 2040 (2). Diabetes and its associated complications are becoming global public health problems and posing a serious challenge in disease management. Many studies have implicated advanced glycation end products (AGEs)1 in the development of insulin resistance, as well as in pathogenesis of diabetic complications (3). The levels of AGEs increase substantially in diabetic plasma due to the hyperglycemic condition. Factors such as oxidative stress, overnutrition, and foods rich in glycating agents promote the formation of AGEs even in nondiabetic condition (4). Oral AGEs foster insulin resistance and diabetes by down-regulation of anti-AGE receptor-1(AGER1), sirtuin 1, and up-regulation of receptor for AGEs (RAGE) (5). AGEs affect glucose uptake, transport and promote insulin resistance in adipocytes (6). While in skeletal muscle cells AGEs inhibit insulin action, mediated through RAGE (7). The AGE-RAGE axis induces oxidative stress, activates proinflammatory pathways and has been considered as a principal pathway in the pathogenesis of diabetes and its complications (8). AGE interacts with RAGE in different cells and tissues, contributing to pathogenesis in diabetes (9). By and large, AGEs contribute to development of insulin resistance leading to diabetes, as well as in the pathogenesis of diabetic complications. Therefore, analysis of plasma AGEs can possibly provide information about the severity of diabetes.Human serum albumin (HSA), one of the most abundant plasma proteins, is highly glycated and contributes predominantly to the plasma AGEs. Apart from its role in pathogenesis, AGE-modified HSA (AGE-HSA) has been suggested as an alternative diagnostic marker to glycated hemoglobin (HbA1c) for monitoring glycemic status in diabetes (10). Although HbA1c is considered the “gold standard” marker, reflecting the glycemic status over the period of 8–10 weeks (1, 10), factors like anemia, blood loss, splenomegaly, and iron deficiency affect HbA1c levels (11). AGE-HSA reflects glycemic status over the preceding 3–4 weeks and has been recommended in gestational diabetes (12). In diabetes, the levels of AGE-HSA increase and were found to be positively correlated with hyperglycemia (13, 14). In addition, several recent studies have suggested that the levels of AGE-HSA are associated with prediabetic condition (15) and microalbuminuria (16). Therefore, quantification of AGE-HSA is of utmost clinical significance. Thus, understanding the site-specific modification and their dynamic transformation to heterogeneous AGEs is quite critical for mass spectrometric quantification.AGEs can be quantified by various approaches, including colorimetric assay, ketoamine oxidase assay, enzyme-linked boronate immunoassay, fluorescence spectroscopy, boronic acid affinity chromatography assay, and mass spectrometry (MS) (17). Among these approaches, MS offers precise characterization of protein glycation, including the amino acid involved in the modification. Most of the AGEs reported in vitro and in vivo were discovered by MS-based techniques (18). AML modification has been extensively studied by different MS approaches. The fragmentation pattern and diagnostic ions for AML rearrangement product has been well established (19, 20). Further specific neutral loss ions of 162 Da, 120 Da, and 84 Da and water loss of 36 Da arising from hexose moiety of glycated peptide were also considered as signature ions to validate the glycation of peptides in HSA (21, 22). Similar characteristic patterns of water loss (18, 36, and 54 Da) ions and immonium ions derived from lysine arising from AML-modified peptide were also used to identify glycated peptides (23, 24). Diagnostic ions serve as the most reliable way of identifying glycated peptide by tandem mass spectrometry. Thus, having a good MS/MS fragment ion is key for precise characterization of glycation. However, the ratio of in vivo AGE-modified to unmodified protein is significantly low, which limits better MS/MS. Therefore, to achieve efficient identification, enrichment of glycated peptides using boronate affinity chromatography (BAC) was adopted prior to MS analysis (25). Further, by using a combination of immunodepletion, enrichment and fractionation strategies, a total of 7,749 unique glycated peptides corresponding to 1,095 native human plasma proteins, 1,592 in vitro glycated human plasma proteins, and 1,664 erythrocyte proteins were identified (26). In these lines, we have previously reported a database search approach for the identification of glycated peptide in a crude or nonenriched sample by untargeted MS/MS or data-independent workflow (27). Glycation is chronic process; a given protein can undergo dynamic heterogeneous transformations as these proteins have varying biological lifespans, influencing the function of a protein. Thus, to assess the degree of glycation at a given pathophysiological condition, precise identification of glycation becomes critical. In this regard, a stable-isotope-dilution tandem mass spectrometry method was employed for simultaneous analysis of CML and CEL in hydrolysates of plasma proteins (28), and 13C6-glucose was utilized to quantify glycated proteins in the plasma and erythrocytes (29, 30). In a recent study, the glycation-sensitive peptides of HSA that could serve as markers for early diagnosis of type 2 diabetes were quantified by using an MS-based 18O-labeling technique (31). However, most of the previous studies have focused on AML modification, rather than other AGE modification. In fact, CML and CEL are the predominant AGEs, constituting up to 80% of total AGEs (32, 33). Diagnostic reporter ions for CML and CEL were reported recently by Prof. Ralf Hoffmann''s group (34). Here, for the first time, we report comprehensive development of an MS/MS fragment ion library for AML, CML, and CEL modifications of albumin. Further, fragment ion library was used as reference for quantification of AML-, CML-, and CEL-modified peptides of albumin in clinical plasma of healthy, prediabetic, diabetic, and microalbuminuria. Targeted SWATH analysis has led to quantification of 13 glycated peptides representing nine lysine sites. These peptides could serve as novel markers in diabetes.  相似文献   

17.
Current strategies to study N-glycoproteins in complex samples are often discrete, focusing on either N-glycans or N-glycosites enriched by sugar-based techniques. In this study we report a simple and rapid sample preparation platform, the GlycoFilter, which allows a comprehensive characterization of N-glycans, N-glycosites, and proteins in a single workflow. Both PNGase F catalyzed de-N-glycosylation and trypsin digestions are accelerated by microwave irradiation and performed sequentially in a single spin filter. Both N-glycans and peptides (including de-N-glycosylated peptides) are separately collected by filtration. The condition to effectively collect complex and heterogeneous N-glycans was established on model glycoproteins, bovine ribonuclease B, bovine fetuin, and human serum IgG. With this platform, the N-glycome, N-glycoproteome and proteome of human urine and plasma were characterized. Overall, a total of 865 and 295 N-glycosites were identified from three pairs of urine and plasma samples, respectively. Many sites were defined unambiguously as partially occupied by the detection of their nonsugar-modified peptides (128 from urine and 61 from plasma), demonstrating that partial occupancy of N-glycosylation occurs frequently. Given the likely high prevalence and variability of partial occupancy, glycoprotein quantification based exclusively on deglycosylated peptides may lead to inaccurate quantification.N-glycosylation is one of the most abundant post-translational modifications of proteins. It is estimated that more than 50% of human proteins are N-glycosylated (1). This type of modification is critical to many fundamental biologic and pathologic processes such as: structural modulation of proteins, cell-cell signaling and interactions, pathogen-host recognition, and tumor progression (2, 3). Inherently, N-glycans are extremely heterogeneous, and subtle variations in the composition or structure may induce dramatic biological consequences (4, 5). Because of their heterogeneity, N-glycans typically need to be released from the parent glycoproteins to be accurately characterized or quantified (3, 6).Identifying the sugar-modified position (glycosite) in a glycoprotein is also critical to understanding the biological role of N-glycosylation (7). Current methods that determine glycosites often use sugar-based enrichment techniques, such as hydrazide chemistry (8) or lectin affinity (9). The extracted glycoproteins or glycopeptides are subjected to de-N-glycosylation, and the deglycosylated peptides are then sequenced by liquid chromatography-tandem MS (LC-MS/MS) to characterize the previously glycosylated sites with a standard bottom-up proteomic approach (10). Filtration has been previously applied to collect peptides (FASP) (11) and deglycosylated peptides after lectin-enrichment (N-glyco FASP) (12). Although these enrichment techniques can identify low-abundant glycosites because of the enrichment selectivity (13), typically they are not feasible for characterization of the N-glycome, because glycans are oxidized and altered when coupled to the hydrazide groups (8), and the selective affinity of the lectin usually biases the N-glycome.N-glycosylation is often considered “irreversible” once a glycoprotein is exported into the extracellular matrix. The addition of a dolichol-linked N-glycan precursor (Glc3Man9GlcNAc2) onto a nascent peptide is an enzyme-catalyzed and nontemplate driven process (2). However, the likelihood and efficiency of this addition are impacted by many factors including: (1) the concentration and activity of oligosaccharyltransferase (2, 14), (2) the availability of dolichol-linked N-glycan precursor (2, 14), (3) the length of time the glycosylated region is unfolded during passage across the endoplasmic reticulum membrane (2, 14), (4) the accessibility of the glycosite, which is greatly impacted by neighboring amino acids (15), and (5) the conformation of a glycosylated protein (correctly folded or not) (14). Because of these variables, the majority of the common N-glycosylation consensus motif (Asn-XXX-Ser/Thr, in which XXX is any amino acid except proline) in human proteins are not actually modified by a sugar chain (15).Furthermore, a glycosite may be partially occupied (PO),1 a state in which both the glycosylated (sugar-modified asparagine) and nonglycosylated (nonsugar-modified asparagine) forms coexist. For example, human corticosteroid-binding globulin, a major plasma glycoprotein with six N-glycosites, has variable degrees in occupancy among its six glycosites (ranging from 70 to 99.5%) that also seem to change with pregnancy (16). Although the biological implications of partial occupancy in N-glycosylation are not well understood, to date, there are no well-defined strategies that can readily identify PO glycosites, particularly in a complex mixture. Current sugar-based enrichment methodologies alone are typically incapable of determining whether a particular glycosite is partially occupied or not, because the nonglycosylated peptides are typically removed.Here we demonstrate a simple, rapid but comprehensive sample preparation platform, the GlycoFilter, which collects N-glycans and peptides separately in a single spin filter device. We demonstrate that glycans, including large acidic glycans, can be effectively separated and captured using a simple shift in pH combined with filtration. Although the lectin-based enrichment method of N-glyco FASP also uses a filtration principle to identify lectin-specific glycosites (12), this platform enables efficient downstream characterization of the N-glycans, N-glycosites, and the remaining proteome of a simple or complex biological sample. Furthermore, the GlycoFilter has the additional nonbiased capability to identify PO N-glycosites using a standard LC-MS/MS approach.  相似文献   

18.
The neuroinvasive pathogen Neisseria meningitidis has 13 capsular serogroups, but the majority of disease is caused by only 5 of these. Groups B, C, Y, and W-135 all display a polymeric sialic acid-containing capsule that provides a means for the bacteria to evade the immune response during infection by mimicking host sialic acid-containing cell surface structures. These capsules in serogroups C, Y, and W-135 can be further acetylated by a sialic acid-specific O-acetyltransferase, a modification that correlates with decreased immunoreactivity and increased virulence. In N. meningitidis serogroup Y, the O-acetylation reaction is catalyzed by the enzyme OatWY, which we show has clear specificity toward the serogroup Y capsule ([Glc-(α1→4)-Sia]n). To understand the underlying molecular basis of this process, we have performed crystallographic analysis of OatWY with bound substrate as well as determined kinetic parameters of the wild type enzyme and active site mutants. The structure of OatWY reveals an intimate homotrimer of left-handed β-helix motifs that frame a deep active site cleft selective for the polysialic acid-bearing substrate. Within the active site, our structural, kinetic, and mutagenesis data support the role of two conserved residues in the catalytic mechanism (His-121 and Trp-145) and further highlight a significant movement of Tyr-171 that blocks the active site of the enzyme in its native form. Collectively, our results reveal the first structural features of a bacterial sialic acid O-acetyltransferase and provide significant new insight into its catalytic mechanism and specificity for the capsular polysaccharide of serogroup Y meningococci.The bacterial pathogen Neisseria meningitidis is a major cause of life-threatening neuroinvasive meningitis in humans (1). In the United States, 75% of bacterial meningitis infections are caused by serogroup C, Y, or W-135 (2). In particular, the proportion of meningococcal infection occurrences in the United States caused by the group Y meningococci has increased significantly from 2% during 1989–1991 to 37% during 1997–2002 (2). Vaccines based on the capsular polysaccharide have been developed for groups A/C/Y/W-135 (2), and the introduction of a group C conjugate vaccine has reduced the incidence and carriage of the C serogroup significantly (3). Although these vaccines are working, they do not yet provide complete protection from meningococcal disease (4).The capsular polysaccharides of N. meningitidis are classified into 13 distinct serogroups based on their chemical structures (5). The capsules of serogroup B and C are homopolymers composed of α-2,8- or α-2,9-linked sialic acid, respectively, whereas serogroup Y and W-135 are heteropolymers of an α-2,6-linked sialic acid on glucose (Y) or galactose (W-135) (6, 7). N. meningitidis group B polysialic acid shares a biochemical epitope with the polysialylated form of the neural cell adhesion molecule of humans (8, 9). Because of this molecular mimicry of the polysialic acid-neural cell adhesion molecule, the bacterial capsular polysaccharide is thus considered a major virulence factor of N. meningitidis (5, 10).Serogroup C, Y, and W-135 of N. meningitidis modify their sialic acid capsules by O-acetylation of the sialic acid (11). Sialic acid is acetylated at the C-7 or C-8 position hydroxyl group in serogroup C, whereas the C-7 or C-9 position is acetylated in serogroup W-135 and Y (11). The O-acetylation of sialic acids is known to alter the physicochemical properties of the polysaccharide capsule (12). In addition, there is growing evidence that O-acetylation of the polysaccharide enhances bacterial pathogenesis by masking the protective epitope in the polysaccharide (1316). For these reasons, considerable effort has been expended to identify and characterize sialic acid O-acetyltransferases in pathogenic bacteria.Recently, the sialic acid-specific O-acetyltransferases from group B Streptococcus, Campylobacter jejuni, Escherichia coli K1, and N. meningitidis serogroup C have been identified (1720) with the latter two variants being the only ones to be characterized biochemically (2123). These studies showed that bacterial sialic acid-specific O-acetyltransferases utilize an acetyl-CoA cofactor as a donor for the acetylation of their capsular sialic acid acceptor substrates (Fig. 1) and identified essential amino acid residues for potential catalytic roles in activity (22, 23). Although the gene encoding the capsule-specific O-acetyltransferase in N. meningitidis serogroup Y (known as OatWY) has been identified, biochemical characterization of the enzyme has not yet been reported. Furthermore, the lack of structural information on a sialic acid O-acetyltransferase from any bacterial species has hampered our ability to further understand the mode of substrate binding, specificity, and catalytic mechanism of this important sialic acid-modifying family.Open in a separate windowFIGURE 1.Reaction scheme of the OatWY-catalyzed O-acetyltransferase. Although acetylation of both the O-7 and O-9 hydroxyl group of the N. meningitidis serogroup Y polysialic acid has been implied through NMR analysis of the corresponding bacterial capsule (11), for simplicity only the O-9 transfer is shown here.Here we report the first kinetic and structural analysis of polysialic acid O-acetyltransferase OatWY from N. meningitidis serogroup Y in complex with either CoA, acetyl-CoA, or S-(2-oxopropyl)-CoA, which is a nonhydrolyzable acetyl-CoA substrate analog. Collectively, this study significantly contributes to our understanding of bacterial polysialic acid O-acetyltransferases, providing valuable insight into how capsular polysaccharide is acetylated in pathogenic bacteria.  相似文献   

19.
Methodology for sequence analysis of ∼150 kDa monoclonal antibodies (mAb), including location of post-translational modifications and disulfide bonds, is described. Limited digestion of fully denatured (reduced and alkylated) antibody was accomplished in seconds by flowing a sample in 8 m urea at a controlled flow rate through a micro column reactor containing immobilized aspergillopepsin I. The resulting product mixture containing 3–9 kDa peptides was then fractionated by capillary column liquid chromatography and analyzed on-line by both electron-transfer dissociation and collisionally activated dissociation mass spectrometry (MS). This approach enabled identification of peptides that cover the complete sequence of a murine mAb. With customized tandem MS and ProSightPC Biomarker search, we verified 95% amino acid residues of this mAb and identified numerous post-translational modifications (oxidized methionine, pyroglutamylation, deamidation of Asn, and several forms of N-linked glycosylation). For disulfide bond location, native mAb is subjected to the same procedure but with longer digestion times controlled by sample flow rate through the micro column reactor. Release of disulfide containing peptides from accessible regions of the folded antibody occurs with short digestion times. Release of those in the interior of the molecule requires longer digestion times. The identity of two peptides connected by a disulfide bond is determined using a combination of electron-transfer dissociation and ion–ion proton transfer chemistry to read the two N-terminal and two C-terminal sequences of the connected peptides.Monoclonal antibodies (mAbs)1 and related biological molecules constitute one of the most rapidly growing classes of human therapeutics. These large proteins (Fig. 1) have molecular weights near 150 kDa and are composed of two identical ∼50 kDa heavy chains (HC) and two identical ∼25 kDa light chains (LC) (1). They also contain at least 16 disulfide bonds that maintain three-dimensional structure and biological activity (2). Although sharing similar secondary protein structures, different mAbs differ greatly in the sequence of variable regions, especially in the complementarity determining regions (CDRs) which are responsible for the diversity and specificity of antibody-antigen binding. Changes to the mAb structure introduced during the manufacturing process or storage may influence the therapeutic efficacy, bio-availability and -clearance, and immunogenic properties and thus alter drug safety (35). Comprehensive characterization of mAbs primary structure, post-translational modifications (PTMs), and disulfide linkages is critical to the evaluation of drug efficacy and safety, as well as understanding the structure/function relationships (4, 6). Presented in this work is a novel protein analytical platform that consists of innovative methods for mass spectrometry (MS) characterization of mAbs. The methodology reported here will have a dramatic impact on the whole field of antibody characterization.Open in a separate windowFig. 1.Diagram of a murine monoclonal antibody structure.Typical MS characterization of proteins uses a “Bottom-Up” approach. This method involves tryptic digestion of the protein(s) into small peptides (mostly below 2500 Da) followed by high-performance liquid chromatography-tandem mass spectrometry (HPLC-MS/MS) analyses of the resulting peptides (7). Although sensitive for MS analysis, small tryptic peptides often have issues such as weak retention in liquid chromatography, difficulties in assigning peptides to specific gene products, and loss of combinatorial PTM information (8). Recent years have seen developments in direct MS analysis of intact proteins (often called “Top-Down” MS). Despite increasing success in characterization of small to medium-sized proteins, MS analysis of intact proteins larger than 50 kDa, including mAbs, is still unsatisfactory because of inefficient gas-phase protein fragmentation and complex fragment ions that restrict efficient data interpretation (9, 10).A “compromise” between Bottom-Up and Top-Down approaches is the “Middle-Down” (or “Middle-Up”) method. Middle-Down analysis typically involves proteolysis using proteases (e.g. Lys-C) or chemicals that hydrolyze proteins at a single type of amino acid residue. This approach aims to generate 3–15 kDa peptides which are compatible with high resolution MS/MS analysis on a chromatographic time scale. The Middle-Down approach inherits some of the advantages of Top-Down analysis, yet has less demanding instrumental requirements compared with intact protein MS in achieving sufficient signal-to-noise ratio (S/N) of fragment ions for sequence mapping (1115).However, limitations of currently available tools for Middle-Down protein analysis are also obvious. First, none of the twenty amino acids is evenly distributed along a polypeptide. Protein digestion at single-type amino acid residues can still produce very small (<1000 Da) or ultra large (>15 kDa) peptides, which deviates from the original intention of the Middle-Down approach (16). Second, the enzymatic digestion efficiency is often low for proteins with highly folded structure or low solubility. Although high concentrations of chaotropic agents such as 8 m urea are often used for protein denaturation, this harsh condition quickly deactivates many commonly used proteases. Third, traditional data-dependent ETD or electron-capture dissociation MS/MS analyses adopt a single reaction parameter for gas-phase dissociation and select only several abundant ions regardless of their charge states. As these methods were previously optimized for tryptic peptide ions that typically carry +2 or +3 charges, they are incompatible with the analysis of large, highly charged peptides that require optimized ETD to achieve high sequence coverage and PTM mapping (12).Herein we report a “time-controlled” proteolysis method for tailored Middle-Down MS analysis of mAb. To hydrolyze the 150 kDa mAb into large peptides for HPLC-MS analysis, we fabricated a capillary enzyme reactor column that contains a specified length of immobilized protease (supplemental Fig. S1 and S2A). Precise control of the sample flow rate leads to defined digestion time of the substrate protein in the reactor. A short digestion time results in a small number of “cuts” along the protein chain and consequently the formation of large peptides (supplemental Fig. S2B). The Bruening group previously demonstrated a similar concept using a nylon membrane electrostatically adsorbed with pepsin or trypsin. Pushing a protein solution (protein dissolved in 5% formic acid solution) through the membrane-based enzyme reactor in less than 1 s breaks the protein into large peptides that facilitate sequence mapping of horse apomyoglobin (17 kDa) and bovine serum albumin (66 kDa) by infusion electrospray ionization MS/MS (17). The advantages of their enzyme disc include simple preparation procedures, as well as the low back pressure in the thin disc that allows for rapid sample flow rate. In our present work, we designed a more robust enzyme reactor that digests alkylated or native mAb into 3–12 kDa peptides in a buffer containing 8 m urea (a condition incompatible with most widely used proteases), and characterized their amino acid sequences, PTMs, as well as the disulfide linkages using HPLC-MS/MS.We chose a rarely used protease, aspergillopepsin I, for the enzyme reactor. Aspergillopepsin I, also known as Aspergillus saitoi acid proteinase, generally catalyzes the hydrolysis of substrate proteins at P1 and P1′ of hydrophobic residues, but also accepts Lys at P1 (18). There are several innovative aspects of employing this enzyme: (1) Aspergillopepsin I is active in 8 m urea at pH 3–4 for at least 1 h. This extreme chaotropic condition may disrupt the higher-order structure of proteins to a great extent and allows for easy access of the protease to most regions of the substrate protein once the disulfide bonds are reduced. (2) Compared with proteases with dual- or single-type amino acid specificity, aspergillopepsin I provides more cleavage sites along an unfolded substrate protein. Allowing limited time for the substrate protein to interact with immobilized aspergillopepsin I should generate large peptides with a relatively narrow size distribution because of similar numbers of missed cleavages on these peptides. (3) The enzyme reactor automatically “quenches” proteolysis as the sample flows out of the column. This is in great contrast to in-tube digestion using solubilized proteases that are active in acidic conditions. In the latter case, digestion is difficult to quench or control because of the sustained enzymatic activity in an acidic condition. (4) Compared with electrostatic or hydrophobic interactions for enzyme immobilization, covalent conjugation of the protease onto porous beads should prevent the replacement of enzymes by upcoming substrate proteins. (5) The enzyme beads can be stored at 4 °C for at least half a year once water is removed, allowing the production of hundreds of disposable enzyme reactors from one batch of beads. In addition, we introduced a new cysteine (Cys) alkylation reagent, N-(2-aminoethyl)maleimide (NAEM) for protein MS analysis. This reagent improves ETD (19) of peptides containing Cys residues by adding a basic, readily protonated side chain to thiol groups.The above features of our new strategy led to the generation of large, highly charged peptides that cover the entire murine mAb. Analyzing ETD and collisionally activated dissociation (CAD) fragments from the most abundant large peptides by ProSightPC revealed near complete sequence coverage of the mAb and multiple PTMs. Furthermore, we digested the native mAb into large fragments of disulfide-bonded peptides using time-controlled digestion. The ETD/ion-ion proton transfer (IIPT) technique (20) allowed facile identification of the N- and C-terminal sequences of two disulfide-bonded peptides and localization of the disulfide bond(s) within/connecting different mAb domains.  相似文献   

20.
Leptospira spp., the causative agents of leptospirosis, adhere to components of the extracellular matrix, a pivotal role for colonization of host tissues during infection. Previously, we and others have shown that Leptospira immunoglobulin-like proteins (Lig) of Leptospira spp. bind to fibronectin, laminin, collagen, and fibrinogen. In this study, we report that Leptospira can be immobilized by human tropoelastin (HTE) or elastin from different tissues, including lung, skin, and blood vessels, and that Lig proteins can bind to HTE or elastin. Moreover, both elastin and HTE bind to the same LigB immunoglobulin-like domains, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12 as demonstrated by enzyme-linked immunosorbent assay (ELISA) and competition ELISAs. The LigB immunoglobulin-like domain binds to the 17th to 27th exons of HTE (17–27HTE) as determined by ELISA (LigBCon4, KD = 0.50 μm; LigBCen7′–8, KD = 0.82 μm; LigBCen9, KD = 1.54 μm; and LigBCen12, KD = 0.73 μm). The interaction of LigBCon4 and 17–27HTE was further confirmed by steady state fluorescence spectroscopy (KD = 0.49 μm) and ITC (KD = 0.54 μm). Furthermore, the binding was enthalpy-driven and affected by environmental pH, indicating it is a charge-charge interaction. The binding affinity of LigBCon4D341N to 17–27HTE was 4.6-fold less than that of wild type LigBCon4. In summary, we show that Lig proteins of Leptospira spp. interact with elastin and HTE, and we conclude this interaction may contribute to Leptospira adhesion to host tissues during infection.Pathogenic Leptospira spp. are spirochetes that cause leptospirosis, a serious infectious disease of people and animals (1, 2). Weil syndrome, the severe form of leptospiral infection, leads to multiorgan damage, including liver failure (jaundice), renal failure (nephritis), pulmonary hemorrhage, meningitis, abortion, and uveitis (3, 4). Furthermore, this disease is not only prevalent in many developing countries, it is reemerging in the United States (3). Although leptospirosis is a serious worldwide zoonotic disease, the pathogenic mechanisms of Leptospira infection remain enigmatic. Recent breakthroughs in applying genetic tools to Leptospira may facilitate studies on the molecular pathogenesis of leptospirosis (58).The attachment of pathogenic Leptospira spp. to host tissues is critical in the early phase of Leptospira infection. Leptospira spp. adhere to host tissues to overcome mechanical defense systems at tissue surfaces and to initiate colonization of specific tissues, such as the lung, kidney, and liver. Leptospira invade hosts tissues through mucous membranes or injured epidermis, coming in contact with subepithelial tissues. Here, certain bacterial outer surface proteins serve as microbial surface components recognizing adhesive matrix molecules (MSCRAMMs)2 to mediate the binding of bacteria to different extracellular matrices (ECMs) of host cells (9). Several leptospiral MSCRAMMs have been identified (1018), and we speculate that more will be identified in the near future.Lig proteins are distributed on the outer surface of pathogenic Leptospira, and the expression of Lig protein is only found in low passage strains (14, 16, 17), probably induced by environmental cues such as osmotic or temperature changes (19). Lig proteins can bind to fibrinogen and a variety of ECMs, including fibronectin (Fn), laminin, and collagen, thereby mediating adhesion to host cells (2023). Lig proteins also constitute good vaccine candidates (2426).Elastin is a component of ECM critical to tissue elasticity and resilience and is abundant in skin, lung, blood vessels, placenta, uterus, and other tissues (2729). Tropoelastin is the soluble precursor of elastin (28). During the major phase of elastogenesis, multiple tropoelastin molecules associate through coacervation (3032). Because of the abundance of elastin or tropoelastin on the surface of host cells, several bacterial MSCRAMMs use elastin and/or tropoelastin to mediate adhesion during the infection process (3335).Because leptospiral infection is known to cause severe pulmonary hemorrhage (36, 37) and abortion (38), we hypothesize that some leptospiral MSCRAMMs may interact with elastin and/or tropoelastin in these elastin-rich tissues. This is the first report that Lig proteins of Leptospira interact with elastin and tropoelastin, and the interactions are mediated by several specific immunoglobulin-like domains of Lig proteins, including LigBCon4, LigBCen7′–8, LigBCen9, and LigBCen12, which bind to the 17th to 27th exons of human tropoelastin (HTE).  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