首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
In the endoplasmic reticulum (ER), lectins and processing enzymes are involved in quality control of newly synthesized proteins for productive folding as well as in the ER-associated degradation (ERAD) of misfolded proteins. ER quality control requires the recognition and modification of the N-linked oligosaccharides attached to glycoproteins. Mannose trimming from the N-glycans plays an important role in targeting of misfolded glycoproteins for ERAD. Recently, two mammalian lectins, OS-9 and XTP3-B, which contain mannose 6-phosphate receptor homology domains, were reported to be involved in ER quality control. Here, we examined the requirement for human OS-9 (hOS-9) lectin activity in degradation of the glycosylated ERAD substrate NHK, a genetic variant of α1-antitrypsin. Using frontal affinity chromatography, we demonstrated that the recombinant hOS-9 mannose 6-phosphate receptor homology domain specifically binds N-glycans lacking the terminal mannose from the C branch in vitro. To examine the specificity of OS-9 recognition of N-glycans in vivo, we modified the oligosaccharide structures on NHK by overexpressing ER α1,2-mannosidase I or EDEM3 and examined the effect of these modifications on NHK degradation in combination with small interfering RNA-mediated knockdown of hOS-9. The ability of hOS-9 to enhance glycoprotein ERAD depended on the N-glycan structures on NHK, consistent with the frontal affinity chromatography results. Thus, we propose a model for mannose trimming and the requirement for hOS-9 lectin activity in glycoprotein ERAD in which N-glycans lacking the terminal mannose from the C branch are recognized by hOS-9 and targeted for degradation.Recognition and sorting of improperly folded proteins is essential to cell survival, and hence, an elaborate quality control system is found in cells. ER4 quality control is well characterized with respect to the N-linked oligosaccharides regulating the folding and degradation of newly synthesized proteins in the ER (1). Immediately after polypeptides enter the ER, Glc3Man9GlcNAc2 (G3M9) precursor oligosaccharides are covalently attached and subsequently processed. Terminally misfolded proteins are removed from the ER by the ERAD machinery (14). Aberrant conformers are recognized, retrotranslocated to the cytosol, and degraded by the ubiquitin-proteasome system (5, 6). Processing of mannose residues from the N-linked oligosaccharides acts as a timer for the recognition of misfolded glycoproteins in the ER lumen (1, 7). ER α1,2-mannosidase I (ER ManI) in mammals and ER α-mannosidase in yeast preferentially trim mannose residues from the middle branch of N-glycans, generating the Man8GlcNAc2 (M8) isomer B (M8B) (8). In mammals, further mannose processing is required as a signal for degradation (1, 9, 10), whereas the presence of M8B is sufficient to signal degradation in yeast (11). The postulated lectin EDEMs in mammals, their yeast homolog Htm1p/Mnl1p, and the yeast MRH domain-containing lectin Yos9p have all been proposed to recognize glycoproteins targeted for degradation (12).The role of Yos9p in glycoprotein ERAD was identified using a genetic screen in Saccharomyces cerevisiae (13). Yos9p, a homolog of hOS-9, contains an MRH domain (14) and functions as a lectin. Yos9p recognizes substrates of the ERAD-lumenal pathway (1517), generating a large ER membrane complex containing the Hrd1p-Hrd3p ubiquitin ligase core complex (1820). The M8B and Man5GlcNAc2 (M5) N-glycans are predicted to function as ligands for Yos9p (17). Bipartite recognition of both glycan and polypeptide by Yos9p has also been reported (15).Recent studies revealed that two mammalian MRH domain-containing lectins, OS-9 and XTP3-B, are ER luminal proteins involved in ER quality control and form a large complex containing the HRD1-SEL1L ubiquitin-ligase in the ER membrane (2124). The components of the complex are similar to yeast, suggesting evolutionary conservation, although the molecular mechanisms underlying the role of OS-9 and XTP3-B remain elusive. Studies using lectin mutants have suggested that the MRH domains are required not for binding to ERAD substrates but for interactions with SEL1L (21), which has multiple N-glycans (25, 26). Additionally, lectin activity appears to be dispensable for hOS-9 binding to misfolded glycoproteins (21, 24). Thus, to understand the role of hOS-9 in the ER quality control pathway, the specific carbohydrate structures recognized by the hOS-9 MRH domain need to be identified, and the requirement of the lectin domain in substrate recognition needs to be determined.In the present study we demonstrate that the lectin activity of hOS-9 is required for enhancement of glycoprotein ERAD. We identified the N-glycan structures recognized by the recombinant hOS-9 MRH domain in vitro by frontal affinity chromatography (FAC). Using a model ERAD substrate, NHK (27), we show that the ability of hOS-9 to enhance ERAD in vivo depends on the oligosaccharides present on NHK, consistent with the FAC results.  相似文献   

3.
Entamoeba histolytica, the protist that causes amebic dysentery and liver abscess, has a truncated Asn-linked glycan (N-glycan) precursor composed of seven sugars (Man5GlcNAc2). Here, we show that glycoproteins with unmodified N-glycans are aggregated and capped on the surface of E. histolytica trophozoites by the antiretroviral lectin cyanovirin-N and then replenished from large intracellular pools. Cyanovirin-N cocaps the Gal/GalNAc adherence lectin, as well as glycoproteins containing O-phosphodiester-linked glycans recognized by an anti-proteophosphoglycan monoclonal antibody. Cyanovirin-N inhibits phagocytosis by E. histolytica trophozoites of mucin-coated beads, a surrogate assay for amebic virulence. For technical reasons, we used the plant lectin concanavalin A rather than cyanovirin-N to enrich secreted and membrane proteins for mass spectrometric identification. E. histolytica glycoproteins with occupied N-glycan sites include Gal/GalNAc lectins, proteases, and 17 previously hypothetical proteins. The latter glycoproteins, as well as 50 previously hypothetical proteins enriched by concanavalin A, may be vaccine targets as they are abundant and unique. In summary, the antiretroviral lectin cyanovirin-N binds to well-known and novel targets on the surface of E. histolytica that are rapidly replenished from large intracellular pools.Entamoeba histolytica causes amebic dysentery and liver abscess in the developing world (10, 20, 29). We are interested in E. histolytica glycoproteins containing Asn-linked glycans (N-glycans) for numerous reasons. E. histolytica makes an N-glycan precursor that contains 7 sugars (Man5GlcNAc2-PP-dolichol) rather than 14 sugars (Glc3Man9GlcNAc2-PP-dolichol) made by most animals, plants, and fungi (21, 31, 44). E. histolytica N-glycans are used for quality control of glycoprotein folding in the endoplasmic reticulum (ER) lumen, and there is positive selection for sites of N-linked glycosylation in secreted and membrane proteins of E. histolytica (5, 11, 53).Unprocessed Man5GlcNAc2, by far the most abundant E. histolytica N-glycan, is present on the plasma membrane and vesicular membranes (31). The antiretroviral lectin cyanovirin-N, which is specific for α-1,2-linked mannose present on unprocessed N-glycans, binds E. histolytica N-glycans and forms aggregates or caps on the surface of E. histolytica trophozoites (1, 25, 31, 44, 45). E. histolytica glycoproteins are also capped by the plant lectin concanavalin A (ConA), which has a broader carbohydrate specificity (mannose and glucose) than cyanovirin-N (3, 16, 18, 19). Heavy subunits of the Gal/GalNAc lectin, the most important E. histolytica vaccine candidate, have 7 to 10 potential sites for N-linked glycosylation (32, 39, 43). Inhibition of N-glycan synthesis results in Gal/GalNAc lectins that are unable to bind to sugars on host epithelial cells.Carbohydrates appear to be an important target on the surface of E. histolytica as anti-proteophosphoglycan (PPG) monoclonal antibodies bind to O-phosphodiester-linked glycans and protect animal models from amebic infection (6, 33, 35, 40, 48). Lectin affinity columns are a powerful method for enriching unique parasite glycoproteins that may be identified by mass spectrometry (MS) of tryptic fragments (17, 55). For example, we recently used the plant lectin wheat germ agglutinin to dramatically enrich glycoproteins with short N-glycans of Giardia (42).The goal of the present studies was to explore further the interaction of the antiretroviral lectin cyanovirin-N with E. histolytica trophozoites in vitro. Questions asked included the following: Are E. histolytica glycoproteins with N-glycans replenished on the plasma membrane after capping with cyanovirin-N? What is the effect of cyanovirin-N capping on other amebic virulence factors and/or vaccine candidates (e.g., the Gal/GalNAc lectin and PPG)? Is capping by cyanovirin-N mediated by actin, as described for capping by the Gal/GalNAc lectin and ConA? What is the effect of the cyanovirin-N on amebic phagocytosis of mucin-coated beads, a surrogate assay for virulence? Which trophozoite glycoproteins are potential targets of cyanovirin-N (identified by mass spectrometry of lectin-enriched E. histolytica proteins)? Are any of them potential vaccine candidates?  相似文献   

4.
5.
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   

6.
7.
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   

8.
9.
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   

10.
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   

11.
Lectin affinity chromatography (LAC) can provide a valuable front-end enrichment strategy for the study of N-glycoproteins and has been used to characterize a broad range eukaryotic N-glycoproteomes. Moreover, studies with mammalian systems have suggested that the use of multiple lectins with different affinities can be particularly effective. A multi-lectin approach has also been reported to provide a significant benefit for the analysis of plant N-glycoproteins; however, it has yet to be determined whether certain lectins, or combinations of lectins are optimal for plant N-glycoproteome profiling; or whether specific lectins show preferential association with particular N-glycosylation sites or N-glycan structures. We describe here a comparative study of three mannose-binding lectins, concanavalin A, snowdrop lectin, and lentil lectin, to profile the N-glycoproteome of mature green stage tomato (Solanum lycopersicum) fruit pericarp. Through coupling lectin affinity chromatography with a shotgun proteomics strategy, we identified 448 putative N-glycoproteins, whereas a parallel lectin affinity chromatography plus hydrophilic interaction chromatography analysis revealed 318 putative N-glycosylation sites on 230 N-glycoproteins, of which 100 overlapped with the shotgun analysis, as well as 17 N-glycan structures. The use of multiple lectins substantially increased N-glycoproteome coverage and although there were no discernible differences in the structures of N-glycans, or the charge, isoelectric point (pI) or hydrophobicity of the glycopeptides that differentially bound to each lectin, differences were observed in the amino acid frequency at the −1 and +1 subsites of the N-glycosylation sites. We also demonstrated an alternative and complementary in planta recombinant expression strategy, followed by affinity MS analysis, to identify the putative N-glycan structures of glycoproteins whose abundance is too low to be readily determined by a shotgun approach, and/or combined with deglycosylation for predicted deamidated sites, using a xyloglucan-specific endoglucanase inhibitor protein as an example.N-glycosylation is one of the most heterogeneous and common post-translational modifications of eukaryotic proteins and one that affects many aspects of protein targeting, enzymatic properties, stability and intermolecular interactions (13). There is therefore considerable interest in developing robust and sensitive high throughput analytical pipelines to isolate and structurally characterize N-glycoprotein populations (2, 48), allowing glycoprotein identification and analysis of the glycosylation site occupancy and N-glycan structure. To this end, lectin affinity chromatography (LAC)1 is increasingly popular: specifically, various lectins are known to have different binding affinities for N-glycans and so the selective binding of N-glycoproteins in complex protein extracts to these lectins and their subsequent release allows a critical enrichment step before sequencing and glycan analysis by MS (9).As a refinement of this approach, the use of multiple lectin affinity chromatography (MLAC) in yeast and animal studies (4, 6, 10), using different proteomic platforms, has been shown to increase the numbers of isolated N-glycoproteins or N-glycopeptides. Collectively, these studies of taxonomically diverse eukaryotic N-glycoproteomes suggest a general conservation of the glycosylation site (N-X-S/T, where X can be any amino acid except proline), as well as conserved features of three-dimensional protein structure (4, 6, 10). Although there have been several studies to determine the structural basis of the binding specificity of specific lectins to yeast and animal N-glycoproteins (11), larger scale N-glycoproteomic analyses have typically not attempted to determine whether a particular combination of lectins provides optimal enrichment, or whether specific features, such as N-glycan structure or amino acid sequence at and around the N-glycosylation site, are associated with different lectins. Therefore, systematic comparative studies are essential to determine whether particular lectins can be optimal for specific tissues, organs, and organisms.LAC has also been used in plant N-glycoprotein analyses to enrich for populations of cell wall localized proteins (8, 12, 13) and a recent report (13) described the application of MLAC to map substantial numbers of N-glycosylation sites in a range of key experimental model organisms, included the plant Arabidopsis thaliana. Using a LTQ-Orbitrap Velos mass spectrometer, the authors identified 2186 unique N-glycosylation sites in proteins extracted from five different arabidopsis organs (13), which represents a substantial increase in the number of identified N-glycosylation sites that have resulted from previous plant N-glycoprotein studies. However, the particular analytical platform that was used did not allow the structural characterization of the N-glycans or N-glycopeptides (13). Indeed, certain features of N-glycopeptides, such as poor fragmentation, heterogeneity and a large dynamic range in most complex mixtures often limits the structural analysis of N-glycans in high throughput systematic analyses (2). It is also important to note that the structures of plant N-glycans differ from those of animals and yeast, as exemplified by the presence of β-1, 2-xylose and α-1, 3-fucose and the complete absence of multiantennary N-glycans and sialic acid in plant N-glycoproteins (1, 14). Therefore, assumptions that are made with regard to the lectin binding of animal and yeast proteins do not necessarily apply to those from plants. Consequently, there is a need to investigate the structural basis of lectin binding to plant N-glycoproteins. Moreover, the limitations of typical shotgun based profiling approaches in identifying and characterizing low abundance N-glycoproteins in complex protein extracts need to be addressed to allow more comprehensive plant N-glycoproteome profiling.In the present study we address both these issues using mature green stage tomato (Solanum lycopersicum) fruit pericarp as an experimental model to carry out a comparative analysis of N-glycoproteins associated with each of three mannose-binding lectins: concanavalin A (ConA), snowdrop lectin (GNA), and lentil lectin (LCH). Fruit development is associated with substantial cell wall metabolism and the expression of many wall localized N-glycoproteins (8) and tomato in particular represents an excellent model for studies of fleshy fruits and cell wall N-glycoproteins (15). We established an MLAC analytical pipeline that included shotgun proteomic profiling and deglycosylation and deamidation analysis, to allow the determination of N-glycoprotein protein identity, N-glycosylation site and N-glycan structures. This information was then used to establish whether a combination of lectins is indeed advantageous for the study of plant N-glycoproteomes, and to assess whether any of these structural characteristics result in predictable preferential binding to specific lectins. From these studies it became evident that large dynamic range of N-glycoprotein abundance was a significant limiting factor in the structural determination of the tomato, and that there was a bias toward the detection of highly abundant N-glycopeptides. We therefore evaluated the use of an in planta recombinant expression strategy, combined with affinity purification MS (AP-MS), as a means to characterize the N-glycan structures of glycoproteins whose abundance is too low to be readily determined via the primary shotgun pipeline, using a tomato xyloglucan-specific endoglucanase inhibitor protein (XEGIP) as a test case.  相似文献   

12.
The accumulation of bioenergy carriers was assessed in two starchless mutants of Chlamydomonas reinhardtii (the sta6 [ADP-glucose pyrophosphorylase] and sta7-10 [isoamylase] mutants), a control strain (CC124), and two complemented strains of the sta7-10 mutant. The results indicate that the genetic blockage of starch synthesis in the sta6 and sta7-10 mutants increases the accumulation of lipids on a cellular basis during nitrogen deprivation relative to that in the CC124 control as determined by conversion to fatty acid methyl esters. However, this increased level of lipid accumulation is energetically insufficient to completely offset the loss of cellular starch that is synthesized by CC124 during nitrogen deprivation. We therefore investigated acetate utilization and O2 evolution to obtain further insights into the physiological adjustments utilized by the two starchless mutants in the absence of starch synthesis. The results demonstrate that both starchless mutants metabolize less acetate and have more severely attenuated levels of photosynthetic O2 evolution than CC124, indicating that a decrease in overall anabolic processes is a significant physiological response in the starchless mutants during nitrogen deprivation. Interestingly, two independent sta7-10:STA7 complemented strains exhibited significantly greater quantities of cellular starch and lipid than CC124 during acclimation to nitrogen deprivation. Moreover, the complemented strains synthesized significant quantities of starch even when cultured in nutrient-replete medium.Microalgae are able to efficiently convert sunlight, water, and CO2 into a variety of products suitable for renewable energy applications, including H2, carbohydrates, and lipids (11, 12, 16, 21, 38, 41, 44). The unicellular green alga Chlamydomonas reinhardtii has emerged as a model organism for studying algal physiology, photosynthesis, metabolism, nutrient stress, and the synthesis of bioenergy carriers (12, 15, 19, 24, 32). During acclimation to nitrogen deprivation, C. reinhardtii cells accumulate significant quantities of starch and form lipid bodies (4, 5, 8, 26, 28, 30, 34, 43, 46, 48). Despite the significance of these products in algal physiology and in biofuels applications, the metabolic, enzymatic, and regulatory mechanisms controlling the partitioning of metabolites into these distinct carbon stores in algae are poorly understood. Several C. reinhardtii starch mutants with various phenotypic changes in starch content and structure have been isolated (2,4). Two of these, the sta6 and sta7 mutants, contain single-gene disruptions that result in “starchless” phenotypes with severely attenuated levels of starch granule accumulation (2, 4, 34, 39, 40, 48).The disrupted loci in the two isolated starchless mutants are distinct and each mutant has a unique phenotype (7, 40). In the sta6 mutant, the small, catalytic subunit of ADP-glucose pyrophosphorylase (AGPase-SS) is disrupted (2, 4, 48), and this mutant accumulates less than 1% of the starch observed in wild-type (WT) cells under conditions of nitrogen deprivation. The sta7 mutant contains a disrupted isoamylase gene (7, 8, 10, 39, 40) and also has severely attenuated levels of starch, but it accumulates a soluble glycogen-like product (4, 9). In this study, we conducted an examination of the unique physiological acclimations that are utilized by these mutants to adapt to the loss of starch synthesis. As the genetic lesions in these two mutants are distinct and block starch synthesis via two very different mechanisms, we investigated the physiological consequences of starch inhibition in both of these mutants from a holistic bioenergy perspective, which included photosynthetic parameters and the overall yields of lipids and carbohydrates, the two primary bioenergy carriers in C. reinhardtii. Specifically, we examined whether the inability to synthesize starch would result in the accumulation of additional lipid, alter cellular growth or cell size, affect acetate utilization, and/or influence photosynthetic O2 evolution. Our data indicate that both the sta6 (BAFJ5) and sta7 (sta7-10) mutants accumulate more lipid than the CC124 control during nitrogen deprivation. However, the additional lipid does not completely offset the loss of starch synthesis from a complete energetic perspective. Increased lipid accumulation during nitrogen stress has also been reported for a variety of starch mutants in recent papers (26, 27, 46). A significant feature in both of the starchless mutants studied here is that O2 evolution and acetate utilization are diminished during nitrogen stress, which is undesirable from an overall bioenergy perspective. Remarkably, complementation of sta7-10 with genomic DNA encoding the wild-type isoamylase gene resulted in cells that were larger than those of the sta6, sta7-10, and CC124 strains, exhibited the highest total lipid levels during nitrogen deprivation, and overaccumulated starch even in nutrient-replete medium.  相似文献   

13.
Various cancers such as colorectal cancer (CRC) are associated with alterations in protein glycosylation. CRC cell lines are frequently used to study these (glyco)biological changes and their mechanisms. However, differences between CRC cell lines with regard to their glycosylation have hitherto been largely neglected. Here, we comprehensively characterized the N-glycan profiles of 25 different CRC cell lines, derived from primary tumors and metastatic sites, in order to investigate their potential as glycobiological tumor model systems and to reveal glycans associated with cell line phenotypes. We applied an optimized, high-throughput membrane-based enzymatic glycan release for small sample amounts. Released glycans were derivatized to stabilize and differentiate between α2,3- and α2,6-linked N-acetylneuraminic acids, followed by N-glycosylation analysis by MALDI-TOF(/TOF)-MS. Our results showed pronounced differences between the N-glycosylation patterns of CRC cell lines. CRC cell line profiles differed from tissue-derived N-glycan profiles with regard to their high-mannose N-glycan content but showed a large overlap for complex type N-glycans, supporting their use as a glycobiological cancer model system. Importantly, we could show that the high-mannose N-glycans did not only occur as intracellular precursors but were also present at the cell surface. The obtained CRC cell line N-glycan features were not clearly correlated with mRNA expression levels of glycosyltransferases, demonstrating the usefulness of performing the structural analysis of glycans. Finally, correlation of CRC cell line glycosylation features with cancer cell markers and phenotypes revealed an association between highly fucosylated glycans and CDX1 and/or villin mRNA expression that both correlate with cell differentiation. Together, our findings provide new insights into CRC-associated glycan changes and setting the basis for more in-depth experiments on glycan function and regulation.Colorectal cancer (CRC)1 is a very prevalent and heterogeneous pathology with highly variable disease progression and clinical outcome among patients. It is the third most common cancer in men and the second most common in women (1) with a highly stage-specific patient survival (2). Treatments are often curative for patients with local disease stages (stage I-II), whereas a 5-year survival of only 13% is observed in patients with distant metastasis (stage IV) (2). As CRC is often asymptomatic in the first years, unfortunately, only 40% of the patients are diagnosed at stage I-II, thus pointing to the urgent need of sensitive diagnostic tools for early detection and consequently effective, curative treatment (3). In this context, understanding the complex mechanisms of CRC is an overriding condition for the development of new, more efficient means of detection, treatment, and prognosis of the disease.Altered glycosylation is a hallmark of cancer (4) and is known to occur with cancer progression (4, 5) as glycans are involved in many cancer-associated events such as adhesion, invasion, and cell signaling (6). As a result of altered glycan structures, cellular processes can be affected due to a change of interactions with glycan-binding proteins (79). Several CRC tissue-associated changes in N-glycans, O-glycans, and glycosphingolipid glycans have been reported and recently reviewed (7). For instance, N-glycans extracted from colorectal tumor tissues are characterized by an increase of sulfated glycans, (truncated) high-mannose-type glycans, and glycans containing sialylated Lewis type epitopes, while showing a decrease of bisection as compared with glycans from nontumor colorectal tissue of the same individuals (10). In accordance, elevated expression of sialyl Lewis A (NeuAcα2,3Galβ1,3[Fucα1,4]GlcNAc-R; NeuAc = N-acetylneuraminic acid, Gal = galactose, Fuc = fucose, GlcNAc = N-acetylglucosamine, R = rest) and pauci-mannosidic N-glycans (truncated high-mannose-type, Man1–4GlcNAc1–4GlcNAc; Man = mannose) was recently found to be correlated with poor prognosis in (advanced) colon carcinomas and N-glycomic profiling was successfully applied to distinguish colorectal adenomas from carcinomas (11).Due to limitations in accessibility of tumor materials and possibilities of in vivo studies on a large scale, cancer cell lines represent a relevant alternative and are widely used as model systems for studying the molecular mechanisms associated with cancer outcome and progression. Since the early 1960s, colorectal cancer cell lines have been established with HT29, LoVo, LS-180, LS-174T, and Co115 representing the first continuous cell lines derived from colon tumors and xenografts (1214). Major benefits of cancer cell lines are their continuous availability, their fast growth, and relatively easy handling, making them suitable also for high-throughput screenings (15) and a large range of experimental possibilities (16). Of note, advantages and limitations of cell lines have been recently reviewed (15).In order to select suitable in vitro models, the characterization of molecular features and their comparison to tumor tissues are needed. A detailed Cancer Cell Line Encyclopedia was recently established containing a genomic dataset for 947 human cancer cell lines, from which 58 are colorectal cancer lineages (17). The Cancer Cell Line Encyclopedia includes data collections on genomic characterization, point mutation frequencies, DNA copy number, and mRNA expression levels. Comparison of these features between cell lines and primary tumors showed a high correlation in most cancer types, especially for colorectal cancer, suggesting that cell lines do represent tumor tissues quite reasonably at least on the genetic level. However, the number of publications characterizing cancer cell lines at a molecular level is far behind the number of articles using cancer cell lines as model systems (18), and only few studies have been conducted on whether in vitro cultured cell lines can serve as suitable models for human tumors (1922). Furthermore, cell lines are well characterized genetically, but they are largely understudied with regard to their glycosylation profiles.Here, we developed and optimized a new analytical method for the more sensitive and higher throughput N-glycome profiling of cells. This method is based on the release of N-glycans in a 96-well plate format from a PVDF-membrane (23) starting from a low number of cells (250,000 cells), the chemical derivatization of released N-glycans enabling the stabilization and discrimination of α2,3- and α2,6-linked N-acetylneuraminic acids (24), followed by registration of the N-glycans by MALDI-TOF(/TOF)-MS. The method was applied to characterize the N-glycome of 25 different colorectal cell lines in a fast and robust manner, including biological and technical replicates for all the cell lines. We obtained the comprehensive N-glycan profiles of 21 cell lines derived from primary tumors, two from lymph node metastases, one from a lung metastasis, and one from ascites fluid to assess their potential as glycobiological tumor model systems. Cancer cell line glycosylation features were then correlated with cancer cell markers and phenotypes as well as glycosyltransferase expressions. This study provides new insights into colon-cancer-associated glycan changes and sets a basis for studies into the functions of N-glycans in CRC with cell lines as model systems.  相似文献   

14.
Glycosylation is one of the most common and important protein modifications in biological systems. Many glycoproteins naturally occur at low abundances, which makes comprehensive analysis extremely difficult. Additionally, glycans are highly heterogeneous, which further complicates analysis in complex samples. Lectin enrichment has been commonly used, but each lectin is inherently specific to one or several carbohydrates, and thus no single or collection of lectin(s) can bind to all glycans. Here we have employed a boronic acid-based chemical method to universally enrich glycopeptides. The reaction between boronic acids and sugars has been extensively investigated, and it is well known that the interaction between boronic acid and diols is one of the strongest reversible covalent bond interactions in an aqueous environment. This strong covalent interaction provides a great opportunity to catch glycopeptides and glycoproteins by boronic acid, whereas the reversible property allows their release without side effects. More importantly, the boronic acid-diol recognition is universal, which provides great capability and potential for comprehensively mapping glycosylation sites in complex biological samples. By combining boronic acid enrichment with PNGase F treatment in heavy-oxygen water and MS, we have identified 816 N-glycosylation sites in 332 yeast proteins, among which 675 sites were well-localized with greater than 99% confidence. The results demonstrated that the boronic acid-based chemical method can effectively enrich glycopeptides for comprehensive analysis of protein glycosylation. A general trend seen within the large data set was that there were fewer glycosylation sites toward the C termini of proteins. Of the 332 glycoproteins identified in yeast, 194 were membrane proteins. Many proteins get glycosylated in the high-mannose N-glycan biosynthetic and GPI anchor biosynthetic pathways. Compared with lectin enrichment, the current method is more cost-efficient, generic, and effective. This method can be extensively applied to different complex samples for the comprehensive analysis of protein glycosylation.Glycosylation is an extremely important protein modification that frequently regulates protein folding, trafficking, and stability. It is also involved in a wide range of cellular events (1) such as immune response (2, 3), cell proliferation (4), cell-cell interactions (5), and signal transduction (6). Aberrant protein glycosylation is believed to have a direct correlation with the development of several diseases, including diabetes, infectious diseases, and cancer (711). Secretory proteins frequently get glycosylated, including those in body fluids such as blood, saliva, and urine (12, 13). Samples containing these proteins can be easily obtained and used for diagnostic and therapeutic purposes. Several glycoproteins have previously been identified as biomarkers, including Her2/Neu in breast cancer (14), prostate-specific antigen (PSA) in prostate cancer (15), and CA125 in ovarian cancer (16, 17), which highlights the clinical importance of identifying glycoproteins as indicators or biomarkers of diseases. Therefore, effective methods for systematic analysis of protein glycosylation are essential to understand the mechanisms of glycobiology, identify drug targets and discover biomarkers.Approximately half of mammalian cell proteins are estimated to be glycosylated at any given time (18). There have been many reports regarding identification of protein glycosylation sites and elucidation of glycan structures (1930). Glycan structure analysis can lead to potential therapeutic and diagnostic applications (31, 32), but it is also critical to identify which proteins are glycosylated as well as the sites at which the modification occurs. Despite progress in recent years, the large-scale analysis of protein glycosylation sites using MS-based proteomics methods is still a challenge. Without an effective enrichment method, the low abundance of glycoproteins prohibits the identification of the majority of sites using the popular intensity-dependent MS sequence method.About a decade ago, a very beautiful and elegant method based on hydrazide chemistry was developed to enrich glycopeptides. Hydrazide conjugated beads reacted with aldehydes formed from the oxidation of cis-diols in glycans (33). This method has been extensively applied to many different types of biological samples (3441). Besides the hydrazide-based enrichment method, lectins have also been frequently used to enrich glycopeptides or glycoproteins before MS analysis (28, 29, 4246). However, there are many different types of lectins, and each is specific to certain glycans (47, 48). Therefore, no combination of lectins can bind to all glycosylated peptides or proteins, which prevents comprehensive analysis of protein glycosylation. Because of the complexity of biological samples, effective enrichment methods are critical for the comprehensive analysis of protein glycosylation before MS analysis.One common feature of all glycoproteins and glycopeptides is that they contain multiple hydroxyl groups in their glycans. From a chemistry point of view, this can be exploited to effectively enrich them. Ideally, chemical enrichment probes must have both strong and specific interactions with multiple hydroxyl groups. The reaction between boronic acids and 1,2- or 1,3-cis-diols in sugars has been extensively studied (4952) and applied for the small-scale analysis of glycoproteins (5355). Furthermore, boronate affinity chromatography has been employed for the analysis of nonenzymatically glycated peptides (56, 57). Boronic acid-based chemical enrichment methods are expected to have great potential for global analysis of glycopeptides when combined with modern MS-based proteomics techniques. However, the method has not yet been used for the comprehensive analysis of protein N-glycosylation in complex biological samples (58).Yeast is an excellent model biological system that has been extensively used in a wide range of experiments. Last year, two papers reported the large-scale analysis of protein N-glycosylation in yeast (59, 60). In one study, a new MS-based method was developed based on N-glycopeptide mass envelopes with a pattern via metabolic incorporation of a defined mixture of N-acetylglucosamine isotopologs into N-glycans. Peptides with the recoded envelopes were specifically targeted for fragmentation, facilitating high confidence site mapping (59). Using this method, 133 N-glycosylation sites were confidently identified in 58 yeast proteins. When combined with an effective enrichment method, this MS-based analysis will provide a more complete coverage of the N-glycoproteome. The other work combined lectin enrichment with digestion by two enzymes (Glu_c and trypsin) to increase the peptide coverage, and 516 well-localized N-glycosylation sites were identified in 214 yeast proteins by MS (60).Here we have comprehensively identified protein N-glycosylation sites in yeast by combining a boronic acid-based chemical enrichment method with MS-based proteomics techniques. Magnetic beads conjugated with boronic acid were systematically optimized to selectively enrich glycosylated peptides from yeast whole cell lysates. The enriched peptides were subsequently treated with Peptide-N4-(N-acetyl-beta-glucosaminyl)asparagine amidase (PNGase F)1 in heavy-oxygen water. Finally, peptides were analyzed by an on-line LC-MS system. Over 800 protein N-glycosylation sites were identified in the yeast proteome, which clearly demonstrates that the boronic acid-based chemical method is an effective enrichment method for large-scale analysis of protein glycosylation by MS.  相似文献   

15.
Despite increasing importance of protein glycosylation, most of the large-scale glycoproteomics have been limited to profiling the sites of N-glycosylation. However, in-depth knowledge of protein glycosylation to uncover functions and their clinical applications requires quantitative glycoproteomics eliciting both peptide and glycan sequences concurrently. Here we describe a novel strategy for the multiplexed quantitative mouse serum glycoproteomics based on a specific chemical ligation, namely, reverse glycoblotting technique, focusing sialic acids and multiple reaction monitoring (MRM). LC-MS/MS analysis of de-glycosylated peptides identified 270 mouse serum peptides (95 glycoproteins) as sialylated glycopeptides, of which 67 glycopeptides were fully characterized by MS/MS analyses in a straightforward manner. We revealed the importance of a fragment ion containing innermost N-acetylglucosamine (GlcNAc) residue as MRM transitions regardless the sequence of the peptides. Versatility of the reverse glycoblotting-assisted MRM assays was demonstrated by quantitative comparison of 25 targeted glycopeptides from 16 proteins between mice with homo and hetero types of diabetes disease model.Clinical proteomics focusing on the identification and validation of biomarkers and the discovery of proteins as therapeutic targets is an emerging and highly important area of proteomics. Biomarkers are measurable indicators of a specific biological state (particularly one relevant to the risk of contraction) and the presence or the stage of disease, and are thus expected to be useful for the prediction, detection, and diagnosis of disease as well as to follow the efficacy, toxicology, and side effects of drug treatment, and to provide new functional insights into biological processes.At present, proteomics methods based on mass spectrometry (MS) have emerged as the preferred strategy for discovery of diagnostic, prognostic, and therapeutic protein biomarkers. Most biomarker discovery studies use unbiased, “identified-based” approaches that rely on high performance mass spectrometers and extensive sample processing. Semiquantitative comparisons of protein relative abundance between disease and control patient samples are used to identify proteins that are differentially expressed and, thus, to populate lists of potential biomarkers. De novo proteomics discovery experiments often result in tens to hundreds of candidate biomarkers that must be subsequently verified in serum. However, despite the large numbers of putative biomarkers, only a small number of them are passed through the development and validation process into clinical practice, and their rate of introduction is declining. The first non-standard abbreviation (MS above is standard) must be footnoted the same as the abbreviation footnote, and MRM must be the first abbreviation in the list because it is the one footnoted. After that the order does not matter.Targeted proteomics using multiple reaction monitoring (MRM)1 is emerging as a technology that complements the discovery capabilities of shotgun strategies as well as an alternative powerful novel MS-based approach to measure a series of candidate biomarkers (17). Therefore, MRM is expected to provide a powerful high throughput platform for biomarker validation, although clinical validation of novel biomarkers has been traditionally relying on immunoassays (8, 9). MRM exploits the unique capabilities of triple quadrupoles (QQQ) MS for quantitative analysis. In MRM, the first and the third quadrupoles act as filters to specifically select predefined m/z values corresponding to the peptide precursor ion and specific fragment ion of the peptide, whereas the second quadrupole serves as collision cell. Several such transitions (precursor/fragment ion pairs) are monitored over time, yielding a set of chromatographic traces with retention time and signal intensity for a specific transition as coordinates. These measurements have been multiplexed to provide 30 or more specific assays in one run. Such methods are slowly gaining acceptance in the clinical laboratory for the routine measurement of endogenous metabolites (10) (e.g. in screening newborns for a panel of inborn errors of metabolism) some drugs (11) (e.g. immunosuppressants), and the component analysis of sugars (12).One of the profound challenges in clinical proteomics is the need to handle highly complex biological mixtures. This complexity presents unique analytical challenges that are further magnified with the use of clinical serum/plasma samples to search for novel biomarkers of human disease. The serum proteome is composed of tens of thousands of unique proteins, of which concentrations may exceed 10 orders of magnitude. Protein glycosylation, one of the most common post-translational modifications, generates tremendous diversity, complexity, and heterogeneity of gene products. It changes the biological and physical properties of proteins, which include functions as signals or ligands to control their distribution, antigenicity, metabolic fate, stability, and solubility. Protein glycosylation, in particular by N-linked glycans, is prevalent in proteins destined for extracellular environments. These include proteins on the extracellular side of the plasma membrane, secreted proteins, and proteins contained in body fluids (such as blood serum, cerebrospinal fluid, urine, breast milk, saliva, lung lavage fluid, or pancreatic juice). Considering that such body fluids are most easily accessible for diagnostic and therapeutic purposes, it is not surprising that many clinical biomarkers and therapeutic targets are glycoproteins. These include, for example, cancer antigen 125 (CA125) in ovarian cancer, human epidermal growth factor receptor 2 (Her2/neu) in breast cancer, and prostate-specific antigen (PSA) in prostate cancer. In addition, changes in the extent of glycosylation and the structure of N-glycans or O-glycans attached to proteins on the cell surface and in body fluids have been shown to correlate with cancer and other disease states, highlighting the clinical importance of this modification as an indicator or effector of pathologic mechanisms (1316). Thus, clinical proteomic platforms should have capability to provide protein glycosylation information as well as sufficient analytical depth to reliably detect and quantify specific proteins with sufficient accuracy and throughput.To improve the detection limits to the required sensitivities, one needs to dramatically reduce the complexity of the sera samples. For focused glycoproteomics, several techniques using lectins or antibodies enabling the large-scale identification of glycoproteins have recently been developed (1719). Notably, Zhang et al. reported a method for the selective isolation of peptides based on chemical oxidation of the carbohydrate moiety and subsequent conjugation to a solid support using hydrazide chemistry (2026). However, it is not possible to provide any structural information about N-glycans because the MS analysis is performed on peptides of which N-glycans are removed preferentially by treating with peptide N-glycanase (PNGase). In 2007, we developed a method for rapid enrichment analysis of peptides bearing sialylated N-glycans on the MALDI-TOF-MS platform (27). The method involves highly selective oxidation of sialic acid residues of glycopeptides to elaborate terminal aldehyde group and subsequent enrichment by chemical ligation with a polymer reagent, namely, reverse glycoblotting technique inspired from an original concept of glycoblotting method (28). This method, in principle, is capable identifying both glycan and peptide sequences concurrently. Recently, Nilsson et al. reported that glycopeptides from human cerebrospinal fluid can be enriched on the basis of the same principle as the reverse glycoblotting protocol, and captured glycopeptides were analyzed with ESI FT-ICR MS (29). Because it is well known that sialic acids play important roles in various biological processes including cell differentiation, immune response, and oncogenesis (3034), our attention has been directed toward feasibility of the reverse glycoblotting technique in quantitative analysis of the specific glycopeptides carrying sialic acid(s) by combining with multiplexed MRM-based MS.  相似文献   

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

18.
Iron is a crucial cofactor in numerous redox-active proteins operating in bioenergetic pathways including respiration and photosynthesis. Cellular iron management is essential to sustain sufficient energy production and minimize oxidative stress. To produce energy for cell growth, the green alga Chlamydomonas reinhardtii possesses the metabolic flexibility to use light and/or carbon sources such as acetate. To investigate the interplay between the iron-deficiency response and growth requirements under distinct trophic conditions, we took a quantitative proteomics approach coupled to innovative hierarchical clustering using different “distance-linkage combinations” and random noise injection. Protein co-expression analyses of the combined data sets revealed insights into cellular responses governing acclimation to iron deprivation and regulation associated with photosynthesis dependent growth. Photoautotrophic growth requirements as well as the iron deficiency induced specific metabolic enzymes and stress related proteins, and yet differences in the set of induced enzymes, proteases, and redox-related polypeptides were evident, implying the establishment of distinct response networks under the different conditions. Moreover, our data clearly support the notion that the iron deficiency response includes a hierarchy for iron allocation within organelles in C. reinhardtii. Importantly, deletion of a bifunctional alcohol and acetaldehyde dehydrogenase (ADH1), which is induced under low iron based on the proteomic data, attenuates the remodeling of the photosynthetic machinery in response to iron deficiency, and at the same time stimulates expression of stress-related proteins such as NDA2, LHCSR3, and PGRL1. This finding provides evidence that the coordinated regulation of bioenergetics pathways and iron deficiency response is sensitive to the cellular and chloroplast metabolic and/or redox status, consistent with systems approach data.The green alga Chlamydomonas reinhardtii has an enormous metabolic versatility (1) and possesses the flexibility to grow in the presence of different carbon sources. It may use carbon dioxide (CO2) for photoautotrophic, acetate for heterotrophic, and both carbon sources for mixotrophic growth. In this alga CO2 is fixed via the Calvin Benson Bassham cycle (2), while acetate can be taken up, converted to acetyl-CoA, and enter the glyoxylate cycle where it may be incorporated into C4 acids (3). In addition to the use of acetate as a source of energy and carbon backbone for biosynthetic processes, acetate can control respiration and photosynthesis in conjunction with the light intensity and CO2 availability (46). Moreover, acclimation responses to iron- and copper-deficiencies significantly vary in photoautotrophic versus heterotrophic conditions (710), indicating that the metabolic status of the cells influence overall cellular acclimation responses.Transition metals like copper, manganese, and iron possess the ability to donate and accept electrons, making these metals suitable cofactors in enzymes that catalyze redox reactions. In particular, iron is used as a cofactor in numerous biochemical pathways and is therefore an essential nutrient. Cells require relatively high levels of iron because it is present in heme-, iron-sulfur and other proteins that function in respiratory and photosynthetic energy transducing. Correspondingly, in eukaryotic cells, the mitochondrion is a major iron-utilizing compartment. It is well established that iron is transported into mitochondria for heme synthesis and iron-sulfur cluster assembly. This is required for the formation of a functional respiratory electron transport machinery (11). Therefore, mitochondrial metabolism in mammals, fungi and plants is significantly affected under iron deficiency, as demonstrated by a number of studies (1214). In plants, the chloroplasts are a primary target of iron deficiency. Changes in chloroplast structure, photosynthetic capacity and the composition of thylakoid membranes have been described for plants deprived of iron (1521).Plants have devised various strategies for acquiring iron (22). Generally, iron deficiency leads to the activation of the iron uptake systems in photosynthetic organisms. For example, the accumulation of the ferroxidase, a component of the high affinity iron uptake system in C. reinhardtii, is very rapidly enhanced when iron becomes limiting (23). Inactivation of IRT1, the most prevalent Fe2+ transporter in Arabidopsis thaliana leads to a dramatic iron deficiency that is reflected by chlorosis (2426). Despite the evolution of elaborate iron-uptake mechanisms in plants, iron deficiency-induced chlorosis remains a major agricultural problem (27, 28).The global impact of iron deficiency on photosynthetic productivity has been also shown in vast ocean regions, which are severely limited for iron (29, 30). Generally, one can conclude that photosynthesis in the oceans and on land can occur in environment where iron availability is restricted.Photosystem I (PSI) is a prime target of iron deficiency as it contains 12 atoms of iron per core complex. In algae, the degradation of PSI is also linked to remodeling of PSI-associated light-harvesting antenna (LHCI) (3133). Cyanobacteria respond to iron deficiency by degradation of light harvesting phycobilisomes (34) and induction of the “iron-stress-induced” gene isiA. The ISIA protein, which has significant sequence similarity with CP43, a chlorophyll a-binding protein of photosystem II (PSII; (35, 36), forms a ring of 18 molecules around a PSI trimeric reaction center, as shown by electron microscopy (37, 38). The overall reorganization of the PSI complex from 900 kDa into 1.7 MDa complex highlights the large adaptive nature of the cellular response to iron deficiency, which helps to optimize the architecture of the photosynthetic apparatus to conditions in which iron is a limiting factor.The marine diatom Thalassiosira oceanica shows a remarkable retrenchment of cellular metabolism and remodeling of bioenergetic pathways in response to iron availability (39). Low iron triggers a reduction in the level of iron-rich photosynthetic proteins while iron-rich mitochondrial proteins are preserved. Furthermore, iron deprivation causes a remodeling of the photosynthetic machinery resulting in the adjustment of light energy use to an overall decline in the level of photosynthetic electron transport complexes (39). These responses, reported for green algae such as C. reinhardtii (31, 40, 41), are important for minimizing photo-oxidative stress and optimizing photosynthetic function. As observed for T. oceanica, under conditions of low iron availability (in the presence of organic carbon) a hierarchy of iron allocation responses in C. reinhardtii result in the down-regulation of iron-rich photosynthetic complexes while iron-rich mitochondrial complexes remain stable (41). Notably, under photoautotrophic and mixotrophic conditions C. reinhardtii displays distinct iron deprivation responses, suggesting that the cell''s response to iron deficiency is also dependent on trophic conditions (79). Thus bioenergetics pathways are remodeled in response to iron availability as well as to the type of carbon source available. Moreover, recent data has indicated that the regulation of iron-induced remodeling of the photosynthetic apparatus is linked to energy metabolism. Depletion of Proton Gradient Regulation Like1 protein (PGRL1) in C. reinhardtii has revealed a decreased efficiency of cyclic electron transfer under low iron conditions resulting in higher vulnerability toward iron deprivation (42).It was our aim to generate a more comprehensive picture of how the proteome of C. reinhardtii varies in response to low iron under distinct trophic conditions and how these changes compare with differences observed for cells grown under photoautotrophic and photoheterotrophic iron replete conditions. Quantitative proteomics in conjunction with a novel hierarchical clustering approach revealed information about the responses of C. reinhardtii to low iron conditions and the iron requirements of photoautotrophic growth. These analyses provide novel insights into the relationships between protein networks required for photosynthesis and iron deprivation-elicited stress responses; these studies are providing the knowledge required for modulating the level of available iron to improve the photosynthetic performance of plants (43, 44).  相似文献   

19.
20.
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