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.
Bacteria/eukaryotes share a common pathway for coenzyme A (CoA) biosynthesis. Although archaeal genomes harbor homologs for most of these enzymes, homologs of bacterial/eukaryotic pantothenate synthetase (PS) and pantothenate kinase (PanK) are missing. PS catalyzes the ATP-dependent condensation of pantoate and β-alanine to produce pantothenate, whereas PanK catalyzes the ATP-dependent phosphorylation of pantothenate to produce 4′-phosphopantothenate. When we examined the cell-free extracts of the hyperthermophilic archaeon Thermococcus kodakaraensis, PanK activity could not be detected. A search for putative kinase-encoding genes widely distributed in Archaea, but not present in bacteria/eukaryotes, led to four candidate genes. Among these genes, TK2141 encoded a protein with relatively low PanK activity. However, higher levels of activity were observed when pantothenate was replaced with pantoate. Vmax values were 7-fold higher toward pantoate, indicating that TK2141 encoded a novel enzyme, pantoate kinase (PoK). A search for genes with a distribution similar to TK2141 led to the identification of TK1686. The protein product catalyzed the ATP-dependent conversion of phosphopantoate and β-alanine to produce 4′-phosphopantothenate and did not exhibit PS activity, indicating that TK1686 also encoded a novel enzyme, phosphopantothenate synthetase (PPS). Although the classic PS/PanK system performs condensation with β-alanine prior to phosphorylation, the PoK/PPS system performs condensation after phosphorylation of pantoate. Gene disruption of TK2141 and TK1686 led to CoA auxotrophy, indicating that both genes are necessary for CoA biosynthesis in T. kodakaraensis. Homologs of both genes are widely distributed among the Archaea, suggesting that the PoK/PPS system represents the pathway for 4′-phosphopantothenate biosynthesis in the Archaea.Coenzyme A (CoA)2 and its derivative 4′-phosphopantetheine are essential cofactors in numerous metabolic pathways, including the tricarboxylic acid cycle, the β-oxidation pathway, and fatty acid and polyketide biosynthesis pathways. Acyl-CoA derivatives are key intermediates in energy metabolism due to their high energy thioester bonds and have been identified in all three domains of life.The mechanism of CoA biosynthesis in bacteria and eukaryotes has been well examined and involves common enzymatic conversions (13). CoA is synthesized from pantothenate via five enzymatic reactions; pantothenate kinase (PanK), 4′-phosphopantothenoylcysteine synthetase (PPCS), 4′-phosphopantothenoylcysteine decarboxylase (PPCDC), 4′- phosphopantetheine adenylyltransferase (PPAT), and dephospho-CoA kinase (DPCK). Although many animals rely on exogenous pantothenate to initiate CoA biosynthesis, microorganisms and plants can synthesize pantothenate from 2-oxoisovalerate and β-alanine. This is a three-step pathway catalyzed by ketopantoate hydroxymethyltransferase (KPHMT), ketopantoate reductase, and pantothenate synthetase (PS).In contrast to the wealth of knowledge on CoA biosynthesis in bacteria and eukaryotes, the corresponding pathway in the Archaea remains unclear (4). Sequence data indicate that the bacterial PPCS and PPCDC homologs and eukaryotic PPAT homologs are found on almost all of the archaeal genomes. The archaeal PPCS and PPCDC genes are fused in many cases, and the bifunctional protein from Methanocaldococcus jannaschii has been shown to exhibit both activities (5). The PPAT homolog from Pyrococcus abyssi has also been studied and confirmed to exhibit the expected PPAT activity (6). Bacterial KPHMT and ketopantoate reductase homologs can also be found, to a lesser extent, on the archaeal genomes. They are not found in the methanogens and Thermoplasmatales, and the fact that the structural similarity among archaeal enzymes is not higher than that toward enzymes from hyperthermophilic bacteria suggests that the archaeal KPHMT and ketopantoate reductase are a result of horizontal gene transfer from bacteria (4). In addition, there are candidate genes distantly related to bacterial/eukaryotic DPCK. However, PS homologs are not found in any of the archaeal genomes, and PanK homologs are found only in a few exceptional cases. Recently, Genschel and co-workers have taken a comparative genomics approach to predict the genes corresponding to the archaeal PS and PanK genes, and have also described the identification of a structurally novel PS from Methanosarcina mazei (4, 7).In this study, we describe the identification of the enzymes responsible for the conversion of pantoate to 4′-phosphopantothenate in Thermococcus kodakaraensis. The organism is a hyperthermophilic archaeon isolated from Kodakara Island, Japan (8, 9). The complete genome sequence is available (10), and gene disruption systems have been developed (1113). To our surprise, the conversion of pantoate to 4′-phosphopantothenate in T. kodakaraensis is not brought about by the two classic enzyme reactions catalyzed by PS and PanK, but by two novel enzyme reactions; phosphorylation of pantoate (pantoate kinase) followed by the condensation of 4-phosphopantoate and β-alanine (4′-phosphopantothenate synthetase or 4-phosphopantoate:β-alanine ligase). Homologs of these two genes are distributed on almost all of the archaeal genomes, suggesting that the Archaea utilize different chemistry in the conversion from pantoate to 4′-phosphopantothenate.  相似文献   

2.
Pantothenate kinase (CoaA) catalyzes the first step of the coenzyme A (CoA) biosynthetic pathway and controls the intracellular concentrations of CoA through feedback inhibition in bacteria. An alternative enzyme found in archaea, pantoate kinase, is missing in the order Thermoplasmatales. The PTO0232 gene from Picrophilus torridus, a thermoacidophilic euryarchaeon, is shown to be a distant homologue of the prokaryotic type I CoaA. The cloned gene clearly complements the poor growth of the temperature-sensitive Escherichia coli CoaA mutant strain ts9, and the recombinant protein expressed in E. coli cells transfers phosphate to pantothenate at pH 5 and 55°C. In contrast to E. coli CoaA, the P. torridus enzyme is refractory to feedback regulation by CoA, indicating that in P. torridus cells the CoA levels are not regulated by the CoaA step. These data suggest the existence of two subtypes within the class of prokaryotic type I CoaAs.Coenzyme A (CoA) is an essential cofactor synthesized from pantothenate (vitamin B5), cysteine, and ATP (1, 20, 30). The thiol group derived from the cysteine moiety in a CoA molecule forms a thioester bond, which is a high-energy bond, with carboxylates including fatty acids. The resulting compounds are called acyl-CoAs (CoA thioesters) and function as the major acyl group carriers in numerous metabolic and energy-yielding pathways. Since it is thought that the pantetheine moiety in CoA existed when life first came about on Earth (25) and at present, a CoA, acyl-CoA, or 4′-phosphopantethein moiety that is common to CoA and acyl carrier proteins is utilized by about 4% of all enzymes as a substrate (6), these compounds are thought to play a crucial role in the earliest metabolic system.Bacteria, fungi, and plants can produce pantothenate, which is the starting material of CoA biosynthesis, although animals must take it from their diet (41). The canonical CoA biosynthetic pathway consists of five enzymatic steps: i.e., pantothenate kinase (CoaA in prokaryotes and PanK in eukaryotes; EC 2.7.1.33), phosphopantothenoylcysteine synthetase (CoaB; EC 6.3.2.5), phosphopantothenoylcysteine decarboxylase (CoaC: EC 4.1.1.36), phosphopantetheine adenylyltransferase (CoaD; EC 2.7.7.3), and dephospho-CoA kinase (CoaE; EC 2.7.1.24). The organisms belonging to the domains Bacteria and Eukarya have this pathway (20, 30). CoaB, CoaC, CoaD, and CoaE are detectable in the complete genome sequences as orthologs of the counterparts from E. coli and humans (15, 16, 32). However, there is diversity among the CoaAs and PanKs, depending on their primary structures, and to date, three types of CoaA in bacteria and one type of PanK in eukaryotes have been identified. CoaAs and PanK catalyze the phosphorylation of pantothenate to produce 4′-phosphopantothenate at the first step of the pathway. First, the Escherichia coli CoaA (CoaAEc) was cloned as a prokaryotic type I CoaA after characterization of the properties enzymatically (42-44, 48). Thereafter, the eukaryotic PanK isoforms were isolated from Aspergillus nidulans (AnPanK), mice (mPanK), and humans (hPanK) (10, 17, 28, 29, 33, 34, 54-56). These enzyme activities were clearly regulated by end products of the biosynthetic pathway such as CoA, acetyl-CoA, and malonyl-CoA, and the pantothenate kinases governed the intracellular concentrations of CoA and acyl-CoAs (10, 17, 28, 29, 33, 34, 43, 44, 48, 54, 55). However, CoaAs insensitive to CoA and acyl-CoAs were recently identified from Staphylococcus aureus (CoaASa), Pseudomonas aeruginosa (CoaAPa), and Helicobacter pylori (CoaAHp) as prokaryotic type II and III CoaAs (9, 11, 18, 27). The structural and functional diversity among pantothenate kinases suggests that they are key indicators of the regulation of the CoA biosynthesis. In archaea neither CoaA nor pantothenate synthetase (PanC; EC 6.3.2.1), which catalyzes the condensation of pantoate and β-alanine to produce pantothenate, had been identified biochemically until very recently. COG1829 and COG1701 were assigned as the respective candidates based on comparative genomic analysis (15). COG1701 was reported to be PanC (36), and later the enzyme was revised to phosphopantothenate synthetase, which catalyzed the condensation of phosphopantoate and β-alanine (52). Together with the identification of COG1701, COG1829 was found to be pantoate kinase, responsible for the phosphorylation of pantoate (52). Homologues of pantoate kinase and phosphopantothenate synthetase are found in most archaeal genomes, thus establishing a noncanonical CoA biosynthetic pathway involving the two novel enzymes. However, homologues of the two novel enzymes are missing in the order Thermoplasmatales.Hence, we proceeded with a search for the kinase genes of the remaining archaea to elucidate the regulatory mechanism(s) underlying archaeal CoA biosynthesis. The PTO0232 gene in the complete genome sequence of Picrophilus torridus was identified as encoding a distant homologue of CoaAEc by a BLAST search. The recombinant protein phosphorylated pantothenate, but the activity was not inhibited at all by CoA or CoA thioesters despite its classification as prokaryotic type I CoaA. This functional difference between P. torridus CoaA (CoaAPt) and CoaAEc can be accounted for by an amino acid substitution at position 247 which possibly interacts with CoA. Here we describe the existence of a second subtype in the class of prokaryotic type I CoaAs.  相似文献   

3.
4.
5.
Cysteine proteases of the papain superfamily are implicated in a number of cellular processes and are important virulence factors in the pathogenesis of parasitic disease. These enzymes have therefore emerged as promising targets for antiparasitic drugs. We report the crystal structures of three major parasite cysteine proteases, cruzain, falcipain-3, and the first reported structure of rhodesain, in complex with a class of potent, small molecule, cysteine protease inhibitors, the vinyl sulfones. These data, in conjunction with comparative inhibition kinetics, provide insight into the molecular mechanisms that drive cysteine protease inhibition by vinyl sulfones, the binding specificity of these important proteases and the potential of vinyl sulfones as antiparasitic drugs.Sleeping sickness (African trypanosomiasis), caused by Trypanosoma brucei, and malaria, caused by Plasmodium falciparum, are significant, parasitic diseases of sub-Saharan Africa (1). Chagas'' disease (South American trypanosomiasis), caused by Trypanosoma cruzi, affects approximately, 16–18 million people in South and Central America. For all three of these protozoan diseases, resistance and toxicity to current therapies makes treatment increasingly problematic, and thus the development of new drugs is an important priority (24).T. cruzi, T. brucei, and P. falciparum produce an array of potential target enzymes implicated in pathogenesis and host cell invasion, including a number of essential and closely related papain-family cysteine proteases (5, 6). Inhibitors of cruzain and rhodesain, major cathepsin L-like papain-family cysteine proteases of T. cruzi and T. brucei rhodesiense (710) display considerable antitrypanosomal activity (11, 12), and some classes have been shown to cure T. cruzi infection in mouse models (11, 13, 14).In P. falciparum, the papain-family cysteine proteases falcipain-2 (FP-2)6 and falcipain-3 (FP-3) are known to catalyze the proteolysis of host hemoglobin, a process that is essential for the development of erythrocytic parasites (1517). Specific inhibitors, targeted to both enzymes, display antiplasmodial activity (18). However, although the abnormal phenotype of FP-2 knock-outs is “rescued” during later stages of trophozoite development (17), FP-3 has proved recalcitrant to gene knock-out (16) suggesting a critical function for this enzyme and underscoring its potential as a drug target.Sequence analyses and substrate profiling identify cruzain, rhodesain, and FP-3 as cathepsin L-like, and several studies describe classes of small molecule inhibitors that target multiple cathepsin L-like cysteine proteases, some with overlapping antiparasitic activity (1922). Among these small molecules, vinyl sulfones have been shown to be effective inhibitors of a number of papain family-like cysteine proteases (19, 2327). Vinyl sulfones have many desirable attributes, including selectivity for cysteine proteases over serine proteases, stable inactivation of the target enzyme, and relative inertness in the absence of the protease target active site (25). This class has also been shown to have desirable pharmacokinetic and safety profiles in rodents, dogs, and primates (28, 29). We have determined the crystal structures of cruzain, rhodesain, and FP-3 bound to vinyl sulfone inhibitors and performed inhibition kinetics for each enzyme. Our results highlight key areas of interaction between proteases and inhibitors. These results help validate the vinyl sulfones as a class of antiparasitic drugs and provide structural insights to facilitate the design or modification of other small molecule inhibitor scaffolds.  相似文献   

6.
The metalloenzyme aminopeptidase P catalyzes the hydrolysis of amino acids from the amino termini of peptides with a prolyl residue in the second position. The human malaria parasite Plasmodium falciparum expresses a homolog of aminopeptidase P during its asexual intraerythrocytic cycle. P. falciparum aminopeptidase P (PfAPP) shares with mammalian cytosolic aminopeptidase P a three-domain, homodimeric organization and is most active with Mn(II) as the cofactor. A distinguishing feature of PfAPP is a 120-amino acid amino-terminal extension that appears to be removed from the mature protein. PfAPP is present in the food vacuole and cytosol of the parasite, a distribution that suggests roles in vacuolar hemoglobin catabolism and cytosolic peptide turnover. To evaluate the plausibility of these putative functions, the stability and kinetic properties of recombinant PfAPP were evaluated at the acidic pH of the food vacuole and at the near-neutral pH of the cytosol. PfAPP exhibited high stability at 37 °C in the pH range 5.0–7.5. In contrast, recombinant human cytosolic APP1 was unstable and formed a high molecular weight aggregate at acidic pH. At both acidic and slightly basic pH values, PfAPP efficiently hydrolyzed the amino-terminal X-Pro bond of the nonapeptide bradykinin and of two globin pentapeptides that are potential in vivo substrates. These results provide support for roles for PfAPP in peptide catabolism in both the food vacuole and the cytosol and suggest that PfAPP has evolved a dual distribution in response to the metabolic needs of the intraerythrocytic parasite.Malaria remains one of the most deadly global infectious diseases with an estimated 500 million clinical cases and 2 million deaths annually (1, 2). Clinical manifestations of the disease arise as the protozoan malaria parasite replicates asexually within human erythrocytes. Five species of the genus Plasmodium infect humans. The cytoadherent properties of red blood cells infected with Plasmodium falciparum, coupled with the ability of the parasite to reach high parasitemia, make it the most virulent species. The emergence of strains of P. falciparum that are resistant to affordable anti-malarial drugs such as chloroquine has complicated efforts to manage malaria, and new drugs are urgently needed.Aminopeptidases catalyze the hydrolysis of amino acids from the amino termini of proteins and peptides. They participate in a wide range of biological processes, including peptide catabolism, protein maturation, antigen presentation on immune cells, and regulation of hormone activity. During the asexual erythrocytic replication cycle of the malaria parasite, aminopeptidases contribute to the catabolism of peptides generated by two major proteolytic pathways. One of these is initiated at the proteasome, a multifunctional protease that plays an important role in the turnover of ubiquitinated cellular proteins in the cytosol (35). In addition, the parasite transports host red blood cell cytosol (consisting primarily of hemoglobin) to an acidic degradative organelle, the food vacuole, where it is degraded in a proteasome-independent pathway (6, 7). As up to 75% of the host cell hemoglobin is catabolized during the intraerythrocytic cycle (8, 9), flux through the vacuolar pathway is substantial. Three aminopeptidases have been identified as key players in recycling amino acids from peptides generated by the proteasomal and vacuolar catabolic pathways: leucine aminopeptidase, aminopeptidase N (PfA-M1), and aminopeptidase P (1014). The latter two enzymes have been found in the food vacuole and therefore may play a direct role in hemoglobin catabolism (11). An aspartyl aminopeptidase is also expressed in asexual stage parasites and hydrolyzes amino-terminal aspartyl and glutamyl substrates (15); however, disruption of its gene does not prevent efficient intraerythrocytic replication (11).Aminopeptidase P (APP)2 homologs exhibit high specificity for proline in the second position of the substrate (the P1′ position in the nomenclature of Schechter and Berger (16)) and catalyze the hydrolysis of the X-Pro amide bond, where X is any aminoacyl residue (17). Because of the cyclic nature of the proline side chain, X-Pro-containing peptides are not easily accommodated in the active sites of broad specificity aminopeptidases (17). In mammals, three APP isozymes have been identified. APP1 is found in the cytosolic fraction of cell lysates and has been characterized from a variety of tissues (1820). Although this enzyme has not, to our knowledge, been localized in intact cells, the apparent lack of specific targeting information is consistent with a role in cytosolic peptide turnover. Active cytosolic forms of APP have been reported in plants (21), fruit flies (22), the microsporidian parasite Encephalitozoon cuniculi (23), and in intestinal cells in Caenorhabditis elegans (24), where it is believed to play a role in the catabolism of peptides produced from ingested bacteria. Mammalian APP2 is a glycosylated ectoenzyme anchored into the membrane of endothelial and epithelial cells with a glycosylphosphatidylinositol attachment (25). The best characterized role of APP2 is the inactivation of the plasma hormone bradykinin, a nonapeptide, through cleavage of the Arg-Pro amino-terminal peptide bond (26, 27). Inhibition of APP2 potentiates the vasodilatory and cardioprotective properties of bradykinin, and APP2 has been considered a target for the development of cardiovascular drugs (2830). A third isoform, APP3, has been identified in the human genome and may be a mitochondrial enzyme but has not yet been characterized (31). Prokaryotic APP homologs contribute to intracellular peptide turnover (32).P. falciparum aminopeptidase P (PfAPP) appears to be important for intraerythrocytic growth, as parasites with a disrupted PfAPP gene could not be isolated (11). We have previously localized a PfAPP-yellow fluorescent protein fusion to the food vacuole and the cytosol of the parasite (11). The cytosolic pool of PfAPP probably fulfills a role in peptide turnover and amino acid recycling that is orthologous to those of the cytosolic enzymes described above. In contrast, there is no report to our knowledge of an aminopeptidase P homolog functioning in an acidic environment akin to the malarial food vacuole. Moreover, characterization of mammalian aminopeptidase P homologs typically reveals a pH optimum of 7–8 with relatively little, if any, activity in the pH range 5.0–5.5 (1820). Although we have previously detected PfAPP activity at acidic pH (11), the catalytic efficiency of the enzyme has not been characterized. Thus, at the outset of this study it was not clear whether PfAPP has a significant catalytic role in the food vacuole.Here we have localized untagged, native PfAPP in the parasite and have confirmed the dual cytosolic/vacuolar distribution of the enzyme. The domain organization, quaternary structure, and metal requirement of PfAPP were characterized. To evaluate the plausibility of a catalytic role for PfAPP at acidic and near-neutral pH, its stability in the pH range 5.0–7.5 was assessed and compared with that of human cytosolic APP1, an enzyme that does not, to our knowledge, have a physiological role in an acidic environment. The catalytic efficiency of PfAPP at a range of pH values was characterized with three X-Pro-containing peptides, two of which are found in the sequences of human α- and β-globin and therefore represent potentially physiological substrates.  相似文献   

7.
Malaria, an infectious disease caused by parasites of the Plasmodium genus, is one of the world''s major public health concerns causing up to a million deaths annually, mostly because of P. falciparum infections. All of the clinical symptoms are associated with the blood stage of the disease, an obligate part of the parasite life cycle, when a form of the parasite called the merozoite recognizes and invades host erythrocytes. During erythrocyte invasion, merozoites are directly exposed to the host humoral immune system making the blood stage of the parasite a conceptually attractive therapeutic target. Progress in the functional and molecular characterization of P. falciparum merozoite proteins, however, has been hampered by the technical challenges associated with expressing these proteins in a biochemically active recombinant form. This challenge is particularly acute for extracellular proteins, which are the likely targets of host antibody responses, because they contain structurally critical post-translational modifications that are not added by some recombinant expression systems. Here, we report the development of a method that uses a mammalian expression system to compile a protein resource containing the entire ectodomains of 42 P. falciparum merozoite secreted and cell surface proteins, many of which have not previously been characterized. Importantly, we are able to recapitulate known biochemical activities by showing that recombinant MSP1-MSP7 and P12-P41 directly interact, and that both recombinant EBA175 and EBA140 can bind human erythrocytes in a sialic acid-dependent manner. Finally, we use sera from malaria-exposed immune adults to profile the relative immunoreactivity of the proteins and show that the majority of the antigens contain conformational (heat-labile) epitopes. We envisage that this resource of recombinant proteins will make a valuable contribution toward a molecular understanding of the blood stage of P. falciparum infections and facilitate the comparative screening of antigens as blood-stage vaccine candidates.Parasites of the Plasmodium genus are the etiological agents responsible for malaria, an infectious disease mostly occurring in developing countries with up to 40% of the world''s population described as being at risk of the disease. Among the Plasmodium species that can affect humans, Plasmodium falciparum is responsible for the highest mortality, causing around one million deaths annually, mostly in children under the age of five (1). The clinical symptoms of malaria occur during the cyclic asexual blood stage of the parasite lifecycle when merozoites, that have invaded and replicated within host erythrocytes, are released into the bloodstream before invading new red blood cells (2). Despite intensive efforts from the research community there is currently no licensed vaccine for malaria. The leading candidate RTS,S/AS01, which targets the pre-erythrocytic stage of the disease and was tested in phase III trials, conferred 30 to 50% protection from clinical malaria, depending on the age group studied (3, 4). This limited efficacy has led to calls for a more effective vaccine and many have suggested that a combinatorial vaccine that additionally targets the blood stage may increase efficacy.A vaccine targeting the proteins expressed on the surface of the blood stage of the parasite is conceptually attractive because merozoites are repeatedly and directly exposed to the human humoral immune system and naturally acquired antibodies against these proteins have been shown to confer at least partial immunity (58). Despite this, only a few antigens discovered before the completion of the parasite genome sequence have been assessed in detail (9) and clinical vaccine trials using antigens that target the blood stage have so far shown limited efficacy, mostly caused by antigenic diversity (10). The sequencing of the parasite genome (11) has identified all possible targets but the systematic screening of these new candidates to assess their potential as a vaccine is hampered by the inability to systematically express recombinant Plasmodium proteins in their native conformation (1215). Likely explanations might be the high (∼80%) A:T content of the P. falciparum genome resulting in low codon usage compatibility in heterologous expression systems, the large size (> 50 kDa) of many proteins, the presence of long stretches of highly repetitive amino acids, and the difficulty in identifying clear structural domains within these proteins using standard prediction computer programs (11). Extracellular proteins, in particular, present an additional challenge because they often have signal peptides and transmembrane regions that can negatively impact expression (1618) and contain structurally important disulfide bonds. However, unlike most other eukaryotic extracellular proteins, Plasmodium cell surface and secreted proteins are not modified by N-linked glycans because of the absence of the necessary enzymes (19).To express Plasmodium proteins for basic research and vaccine development, a diverse range of expression systems have been tried (12) ranging from bacteria (17, 18), yeast (13), Dictyostelium (20), and plants (21) to mammalian cells (22) and cell-free systems (2325). To circumvent the problem of codon usage, bacterial (26) and yeast (27) strains with modified tRNA pools have been developed, or sequences of the gene of interest synthesized and codon-optimized to match that of the expression host (28, 29). Although Escherichia coli has been the most popular expression system because of its relative simplicity and cost effectiveness, large-scale production of soluble functional Plasmodium falciparum recombinant proteins remains challenging with success rates ranging from just 6 to 21% (17, 18) and is often hindered by the need for complex refolding procedures. Similarly, attempts have been made to compile large panels of parasite proteins using in vitro translation systems (23, 25, 30, 31). These systems, however, require reducing conditions and are therefore not generally suitable for the systematic expression of extracellular proteins that occupy an oxidizing environment and critically require the formation of disulfide bonds for proper function. As a result, functional analyses of extracellular parasite proteins have often been restricted to smaller subfragments of the proteins that can be expressed in a soluble form rather than the entire extracellular region. Although eukaryotic expression systems are able to add disulfide bonds, they also often inappropriately glycosylate parasite proteins, adding further complication (32). A generic method that would overcome these technical challenges to express, in a systematic way, panels of recombinant Plasmodium proteins that have retained their native function and conformation would therefore be a valuable resource for the molecular investigations of erythrocyte invasion and the development of a blood stage vaccine.To generate a resource of correctly folded recombinant merozoite proteins, we used a mammalian expression system and established the parameters necessary for high-level expression. Using this method, we compiled a panel of 42 proteins that corresponds to the repertoire of abundant cell surface and secreted merozoite proteins of the 3D7 strain of Plasmodium falciparum. Biochemical activity of these proteins was demonstrated by recapitulating known protein interactions and by showing conformation-sensitive immunoreactivity of the recombinant proteins using immune sera.  相似文献   

8.
9.
The leading malaria vaccine candidate, RTS,S, targets the sporozoite and liver stages of the Plasmodium falciparum life cycle, yet it provides partial protection against disease associated with the subsequent blood stage of infection. Antibodies against the vaccine target, the circumsporozoite protein, have not shown sufficient correlation with risk of clinical malaria to serve as a surrogate for protection. The mechanism by which a vaccine that targets the asymptomatic sporozoite and liver stages protects against disease caused by blood-stage parasites remains unclear. We hypothesized that vaccination with RTS,S protects from blood-stage disease by reducing the number of parasites emerging from the liver, leading to prolonged exposure to subclinical levels of blood-stage parasites that go undetected and untreated, which in turn boosts pre-existing antibody-mediated blood-stage immunity. To test this hypothesis, we compared antibody responses to 824 P. falciparum antigens by protein array in Mozambican children 6 months after receiving a full course of RTS,S (n = 291) versus comparator vaccine (n = 297) in a Phase IIb trial. Moreover, we used a nested case-control design to compare antibody responses of children who did or did not experience febrile malaria. Unexpectedly, we found that the breadth and magnitude of the antibody response to both liver and asexual blood-stage antigens was significantly lower in RTS,S vaccinees, with the exception of only four antigens, including the RTS,S circumsporozoite antigen. Contrary to our initial hypothesis, these findings suggest that RTS,S confers protection against clinical malaria by blocking sporozoite invasion of hepatocytes, thereby reducing exposure to the blood-stage parasites that cause disease. We also found that antibody profiles 6 months after vaccination did not distinguish protected and susceptible children during the subsequent 12-month follow-up period but were strongly associated with exposure. Together, these data provide insight into the mechanism by which RTS,S protects from malaria.The RTS,S malaria vaccine candidate provides partial protection against clinical malaria in African children, which has been repeatedly demonstrated in Phase IIb and Phase III clinical trials (15). The RTS,S target is the Plasmodium falciparum circumsporozoite protein (CSP), and it has been shown to generate high antibody titers that remain above levels acquired naturally for years (6). However, it remains unclear how the vaccine, which targets sporozoites, provides protection against disease caused by blood-stage parasites. A rational mechanism has been proposed, based on antibody and T cell responses to the CSP (7), but antibodies have not consistently correlated with protection when clinical disease was the trial end point (8). We and others hypothesized that partial blockage of pre-erythrocytic development would result in low-level blood-stage infections that go untreated in RTS,S vaccinees and that this would boost the blood-stage immune response, contributing to protection from malaria disease (810).We set out to address the question of how the vaccine works by investigating the response to malaria parasites in the context of RTS,S vaccination. However, until recently, the means of assessing the response to malaria parasites has been limited to a sparse selection of recombinant proteins or parasite lysates. The P. falciparum (Pf) proteome contains more than 5,300 proteins, and, until recently, less than 0.5% of them have been closely investigated (11). Similar to the approach taken with gene expression microarrays, protein arrays offer the opportunity to screen antibody responses to partial or complete proteomes (12). This approach was taken in this study to identify the breadth and magnitude of naturally acquired immune responses in Mozambican children vaccinated with RTS,S/AS021, the predecessor to the RTS,S/AS01 formulation used in the current Phase III trial, or comparator vaccine.In addition to characterizing the RTS,S mode of action, we aimed to identify biomarker correlates of protection against clinical malaria. Malaria vaccinology is lacking in surrogate markers of protection, and such biomarkers would be a highly useful measure for assessment of vaccine efficacy, especially when control or placebo vaccine groups are no longer available (13). This could mitigate the current inefficient means of measuring efficacy in clinical trials. In the post-genomic era, with systems approaches employed for questions to complex problems in biology and medicine, perhaps alternative thinking is required to tackle the question of how to assess vaccines (14, 15). In this study, we took steps in that direction in order to identify antibody signatures of protection that contribute toward a surrogate marker for the RTS,S and other vaccines.  相似文献   

10.
Plants make coenzyme A (CoA) in the cytoplasm but use it for reactions in mitochondria, chloroplasts, and peroxisomes, implying that these organelles have CoA transporters. A plant peroxisomal CoA transporter is already known, but plant mitochondrial or chloroplastic CoA transporters are not. Mitochondrial CoA transporters belonging to the mitochondrial carrier family, however, have been identified in yeast (Saccharomyces cerevisiae; Leu-5p) and mammals (SLC25A42). Comparative genomic analysis indicated that angiosperms have two distinct homologs of these mitochondrial CoA transporters, whereas nonflowering plants have only one. The homologs from maize (Zea mays; GRMZM2G161299 and GRMZM2G420119) and Arabidopsis (Arabidopsis thaliana; At1g14560 and At4g26180) all complemented the growth defect of the yeast leu5Δ mitochondrial CoA carrier mutant and substantially restored its mitochondrial CoA level, confirming that these proteins have CoA transport activity. Dual-import assays with purified pea (Pisum sativum) mitochondria and chloroplasts, and subcellular localization of green fluorescent protein fusions in transiently transformed tobacco (Nicotiana tabacum) Bright Yellow-2 cells, showed that the maize and Arabidopsis proteins are targeted to mitochondria. Consistent with the ubiquitous importance of CoA, the maize and Arabidopsis mitochondrial CoA transporter genes are expressed at similar levels throughout the plant. These data show that representatives of both monocotyledons and eudicotyledons have twin, mitochondrially located mitochondrial carrier family carriers for CoA. The highly conserved nature of these carriers makes possible their reliable annotation in other angiosperm genomes.CoA acts as an acyl carrier in many reactions of primary and secondary metabolism, and some 8% of the nearly 4,900 enzymes described in the Enzyme Commission database are CoA dependent (Bairoch, 2000). CoA occupies a central position in lipid metabolism, respiration, gluconeogenesis, and other pathways (Leonardi et al., 2005). It is present in all forms of life, but while all organisms can synthesize it from pantothenate (vitamin B5), only prokaryotes, plants, and fungi are able to synthesize pantothenate; animals obtain pantothenate from the diet (Daugherty et al., 2002; Leonardi et al., 2005; Webb and Smith, 2011).In plants, the steps that convert pantothenate to CoA are almost certainly cytosolic (Webb and Smith, 2011; Gerdes et al., 2012). CoA, however, is required in mitochondria for the citric acid cycle, in chloroplasts for fatty acid synthesis, and in peroxisomes for β-oxidation. CoA, therefore, must be imported into these organelles from the cytosol, and indeed, early work demonstrated a CoA transport system in potato (Solanum tuberosum) mitochondria (Neuburger et al., 1984). Yeast (Saccharomyces cerevisiae) and mammalian mitochondria and peroxisomes likewise import CoA because they cannot make it (Fiermonte et al., 2009; Agrimi et al., 2012b). The compartmentation of CoA in all eukaryotes appears to be closely regulated, with cytosol and organelles maintaining separate CoA pools whose levels can modulate fluxes through CoA-dependent reactions (Hunt and Alexson, 2002; Leonardi et al., 2005; De Marcos Lousa et al., 2013).Mitochondrial CoA transporters belonging to the mitochondrial carrier family (MCF) have been identified in yeast (Leu-5p; Prohl et al., 2001) and human (SLC25A42; Fiermonte et al., 2009). Furthermore, peroxisomal CoA carriers from human (SLC25A17; Agrimi et al., 2012b) and Arabidopsis (Arabidopsis thaliana; peroxisomal CoA and NAD carrier [PXN]; Agrimi et al., 2012a) have also been identified. However, no transporters for CoA are known for plant mitochondria or chloroplasts (Palmieri et al., 2011; Gerdes et al., 2012).In this study, a comparative genomic analysis first identified close Arabidopsis and maize (Zea mays) homologs of the yeast and mammalian mitochondrial CoA carriers as candidates for the missing plant mitochondrial or chloroplast transporters. Experimental evidence then demonstrated that the candidate proteins transport CoA when expressed in yeast, that they are targeted to mitochondria in vitro and in planta, and that they are expressed throughout the plant.  相似文献   

11.
Sphingosine 1-phosphate (S1P) is a bioactive lipid signal transmitter present in blood. Blood plasma S1P is supplied from erythrocytes and plays an important role in lymphocyte egress from lymphoid organs. However, the S1P export mechanism from erythrocytes to blood plasma is not well defined. To elucidate the mechanism of S1P export from erythrocytes, we performed the enzymatic characterization of S1P transporter in rat erythrocytes. Rat erythrocytes constitutively released S1P without any stimulus. The S1P release was reduced by an ABCA1 transporter inhibitor, glyburide, but not by a multidrug resistance-associated protein inhibitor, MK571, or a multidrug resistance protein inhibitor, cyclosporine A. Furthermore, we measured S1P transport activity using rat erythrocyte inside-out membrane vesicles (IOVs). Although the effective S1P transport into IOVs was observed in the presence of ATP, this activity was also supported by dATP and adenosine 5′-(β,γ-imido)triphosphate. The rate of S1P transport increased depending on S1P concentration, with an apparent Km value of 21 μm. Two phosphorylated sphingolipids, dihydrosphingosine 1-phosphate and ceramide 1-phosphate, did not inhibit S1P transport. Similar to the intact erythrocytes, the uptake of S1P into IOVs was inhibited by glyburide and vanadate but not by the other ABC transporter inhibitors. These results suggest that S1P is exported from the erythrocytes by a novel ATP-dependent transporter.Sphingosine 1-phosphate (S1P),2 a bioactive lipid molecule present in the blood, plays an important role in diverse cellular responses, such as migration, proliferation, and differentiation (1, 2). These processes are triggered by the binding of S1P to its specific receptors (3), of which five subtypes (S1P1-S1P5) have been identified in endothelial and immune cells (4). Studies using S1P1 receptor-deficient mice showed abnormalities in lymphocyte egress from lymph nodes, spleen, and thymus (5, 6). Whereas blood plasma contains a basal level of S1P from the nanomolar to the micromolar range (712), lymphoid tissues maintain a low S1P environment through the activity of S1P lyase (13). It has been proposed that a higher concentration of S1P in the blood plasma than in the lymphoid organs establishes an essential gradient along which lymphocytes expressing the S1P1 receptor on cell surfaces migrate (2, 5, 6, 1315).The source of plasma S1P remains unclear despite its importance in the cellular responses of endothelial cells and lymphocytes. Unlike most cells, blood cells, astrocytes, and vascular endothelial cells are reported to release S1P (8, 1618). These cells contain sphingosine kinase, which synthesizes S1P through the phosphorylation of sphingosine (16, 18, 19). Whereas platelets and mast cells release S1P in a stimulus-dependent manner (17, 20), erythrocytes, neutrophils, and mononuclear cells release S1P in a stimulus-independent manner (16). The roles of S1P derived from erythrocytes, the most abundant of these blood cells, have not been elucidated. However, recent reports suggest that S1P released from erythrocytes is a major source of plasma S1P (7, 9) and promotes lymphocyte egress to blood (9).Previously, we showed that S1P is released from rat platelets upon stimulation by thrombin or Ca2+ (21). We proposed that an ATP-dependent transporter plays a key role in S1P release from platelets (21). However, the detailed mechanism of S1P release is unclear because there is no way to assay the transport of S1P across the membrane. In this study we compared the properties of S1P release from erythrocytes with that of platelets and showed that S1P release from erythrocytes does not require any stimuli. We then established an assay to measure the ATP-dependent S1P uptake into inside-out membrane vesicles (IOVs) prepared from rat erythrocytes and characterized S1P transport in erythrocytes.  相似文献   

12.
Glycosaminoglycan (GAG) biosynthesis requires numerous biosynthetic enzymes and activated sulfate and sugar donors. Although the sequence of biosynthetic events is resolved using reconstituted systems, little is known about the emergence of cell-specific GAG chains (heparan sulfate, chondroitin sulfate, and dermatan sulfate) with distinct sulfation patterns. We have utilized a library of click-xylosides that have various aglycones to decipher the mechanism of GAG biosynthesis in a cellular system. Earlier studies have shown that both the concentration of the primers and the structure of the aglycone moieties can affect the composition of the newly synthesized GAG chains. However, it is largely unknown whether structural features of aglycone affect the extent of sulfation, sulfation pattern, disaccharide composition, and chain length of GAG chains. In this study, we show that aglycones can switch not only the type of GAG chains, but also their fine structures. Our findings provide suggestive evidence for the presence of GAGOSOMES that have different combinations of enzymes and their isoforms regulating the synthesis of cell-specific combinatorial structures. We surmise that click-xylosides are differentially recognized by the GAGOSOMES to generate distinct GAG structures as observed in this study. These novel click-xylosides offer new avenues to profile the cell-specific GAG chains, elucidate the mechanism of GAG biosynthesis, and to decipher the biological actions of GAG chains in model organisms.Proteoglycans play a major role in various cellular/physiological processes, including blood clotting, growth factor signaling, embryogenesis, axon growth and guidance, angiogenesis, and others (14). Proteoglycans consists of a core protein and glycosaminoglycan (GAG)2 chains. GAG chains account for >50% of the total molecular weight and are primarily responsible for physiological activity of the proteoglycans (5, 6). GAG chains are composed of repeating disaccharide units of a hexosamine residue and a hexuronic acid residue. The three major types of GAG chains found in the proteoglycans are heparan sulfate (HS), chondroitin sulfate (CS) and dermatan sulfate (DS). These GAG chains are differentiated by the type of hexosamine (glucosamine/galactosamine), the percentage of uronic acid epimers (glucuronic/iduronic acid), the extent of sulfation, and the nature of glycosidic linkage (α-/β-). One of the key steps in the proteoglycan biosynthesis is the xylosylation of certain specific serine residues of the core protein (710), which occurs in the late endoplasmic reticulum and/or cis-Golgi compartments (1113). This key event is an essential prelude for the construction of the proteoglycan linkage region (14) that is followed by sequence of events resulting in the assembly of mature GAG chains by alternative addition of hexosamine and glucuronic acid residues. The maturation of GAG chains occurs in the medial and trans-Golgi compartments and involves the following events: N-sulfation of glucosamine units by N-deacetylase-N-sulfotransferases (for HS only), epimerization of glucuronic acids to iduronic acids by C-5 epimerase, and sulfation of the repeating disaccharide units by a variety of sulfotransferases and their isoforms.The position, extent, and pattern of sulfation attribute enormous diversity to GAG chains, which confer specificity in binding to a vast array of proteins. These diverse structural features are very tightly regulated in a spatio-temporal manner during and beyond the development of an organism, and these features dictate differential interactions with various growth factors and receptors, and numerous protein targets leading to an array of physiological functions (15, 16).The presence of free GAG chains has been known to disrupt the interaction of endogenous GAG components of proteoglycans with protein ligands thereby altering the physiological activities. Consequently, they have been used as molecular tools in the elucidation of the role of GAG chains in the activation of cellular events (1719). Free GAG chains can be synthesized in vitro in cell culture by providing exogenous xylosides containing various hydrophobic aglycone moieties. Thus, the xylosides can act as false acceptors for initiation of linkage region and the subsequent elongation of GAG chains. Xylosides have been used for over three decades both in vitro (2028) and in vivo (25, 2931) to probe the functional significance of GAG chains in various dynamic systems under different conditions. The quantity and type of GAG chains synthesized depends on the system where it was tested and on the structure of the aglycone moiety of the xylosides (3234). Most of these studies have utilized a few O-xylosides that are inherently less stable. Furthermore, synthesis of O-xylosides requires very stringent reaction conditions, toxic Lewis acids, and at times leads to inseparable α and β mixtures with unpredictable yields. As a result, it is tedious to generate diverse xylosides in a rapid fashion and utilize them in biological systems. We envisioned that synthesis of metabolically stable xylosides will advance our knowledge of glycosaminoglycan biosynthesis and how they regulate various pathophysiological processes.In our earlier communication, we outlined a simple strategy, utilizing click chemical methodology that addresses the above limitations of O-xylosides, to generate a library of xylosides in a robust manner (35). Several studies have shown that the concentration of the primers and the aglycone moieties influence the composition of GAG chains produced (32). In the current study, we show that the aglycone moieties of click-xylosides may not only influence the composition and quantity of GAG chains but also the extent of sulfation, sulfation pattern, disaccharide composition, and chain length using pgsA-745 Chinese hamster ovary (CHO) cell line as a model cellular system. Our findings provide new insights in to the mechanism of GAG biosynthesis and offer new avenues to decipher the biological actions of GAG chains in model organisms.  相似文献   

13.
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   

14.
15.
16.
Intracellular pathogens contribute to a significant proportion of infectious diseases worldwide. The successful strategy of evading the immune system by hiding inside host cells is common to all the microorganism classes, which exploit membrane microdomains, enriched in cholesterol and sphingolipids, to invade and colonize the host cell. These assemblies, with distinct biochemical properties, can be isolated by means of flotation in sucrose density gradient centrifugation because they are insoluble in nonionic detergents at low temperature. We analyzed the protein and lipid contents of detergent-resistant membranes from erythrocytes infected by Plasmodium falciparum, the most deadly human malaria parasite. Proteins associated with membrane microdomains of trophic parasite blood stages (trophozoites) include an abundance of chaperones, molecules involved in vesicular trafficking, and enzymes implicated in host hemoglobin degradation. About 60% of the identified proteins contain a predicted localization signal suggesting a role of membrane microdomains in protein sorting/trafficking.To validate our proteomic data, we raised antibodies against six Plasmodium proteins not characterized previously. All the selected candidates were recovered in floating low-density fractions after density gradient centrifugation. The analyzed proteins localized either to internal organelles, such as the mitochondrion and the endoplasmic reticulum, or to exported membrane structures, the parasitophorous vacuole membrane and Maurer''s clefts, implicated in targeting parasite proteins to the host erythrocyte cytosol or surface. The relative abundance of cholesterol and phospholipid species varies in gradient fractions containing detergent-resistant membranes, suggesting heterogeneity in the lipid composition of the isolated microdomain population. This study is the first report showing the presence of cholesterol-rich microdomains with distinct properties and subcellular localization in trophic stages of Plasmodium falciparum.Plasmodium falciparum, the most deadly agent of human malaria, caused around 216 million infections and 655,000 deaths in 2010. The complex parasite life cycle involves the development in a mosquito vector of the Anopheles genus and eventual migration to a human host. In this host, asymptomatic multiplication in the liver cells is followed by parasite release into the bloodstream and erythrocyte invasion. Inside the erythrocytes, parasites grow (trophozoite stage) and multiply asexually (schizont stage), developing into highly specialized invasive forms (merozoites). A fraction of parasites differentiate into gametocytes, the gamete precursors necessary to complete the transmission cycle. Parasite blood stages, responsible for malaria pathogenesis and transmission, actively remodel the host erythrocyte, generating novel membrane compartments to sustain the export and sorting of proteins to the host cell cytosol, membrane skeleton, and plasma membrane. The parasitophorous vacuole membrane (PVM),1 which surrounds the parasite throughout the erythrocytic cycle, is the site where exported proteins are translocated into the erythrocyte cytosol (1, 2). Membrane-bound structures of parasite origin, the so-called Maurer''s clefts (MCs) (3, 4), form functionally independent compartments at the red blood cell (RBC) periphery and mediate the sorting/assembly of virulence factors en route to the host cell surface (5). In addition, populations of different vesicles (25 and 80 nm) were identified in the RBC cytosol, suggesting the presence of vesicular mediated trafficking for the delivery of cargo to different destinations (6).Membranes are important sites for cellular signaling events, and many proteins with therapeutic potential localize in these cellular compartments (7, 8). Membrane microdomains enriched in sphingolipids and cholesterol, also referred to as lipid rafts, have been extensively studied in different cell types and gained particular interest for their roles in infection and pathogenesis (8, 9). These assemblies are small and dynamic and can be stabilized to form larger microdomains implicated in a wide range of fundamental cellular processes, which vary depending on cell type (10). Sphingolipids exhibit strong lateral cohesion, generating tightly packed regions in the membrane bilayer, and cholesterol acts as a spacer present in both membrane leaflets generating stable, liquid-ordered phase domains in the membrane bilayer (11). Distinct biochemical properties render these membrane assemblies insoluble in nonionic detergents at low temperature, allowing for their enrichment as detergent-resistant membranes (DRMs). Proteins with DRM-raft affinity include glycosylphosphatidyl inositol (GPI)-anchored proteins and acylated, myristoylated, and palmitoylated proteins (11). DRM rafts also restrict free diffusion of membrane proteins, thereby directing the trafficking of proteins and lipids to and from cellular compartments. Because of their endocytic and receptor clustering capacity, an increasing number of pathogens, including Plasmodium falciparum, utilize them when interacting with their target cells for invasion (9, 12).Even though cholesterol-rich membrane microdomains are implicated in fundamental processes in the parasite life cycle, Plasmodium is unable to synthesize sterols and depends entirely on hosts for its cholesterol supply. During merozoite invasion, lipid and protein components of the erythrocyte rafts are selectively recruited and incorporated into the nascent PVM (13, 14). Plasmodium liver stages utilize cholesterol internalized by low-density lipoprotein and synthesized by hepatocytes (15).To shed light on the organization and dynamics of these assemblies during parasite development inside the infected cell, we identified and validated the DRM-raft proteome of the P. falciparum trophozoite/early schizont. Detected proteins only partially overlap with DRM components of the P. falciparum late schizonts (16, 17) or the mixed blood stages of the rodent malaria agent P. berghei (18). Immunolocalization of selected DRM-associated proteins indicated that these assemblies may reside in both exported compartments (PVM, MCs) and intracellular membranes/organelles. The analysis of DRM lipids suggested that distinct microdomains exist in the infected erythrocyte that differ in their relative abundance of cholesterol and phospholipids.  相似文献   

17.
18.
Malaria parasites actively remodel the infected red blood cell (irbc) by exporting proteins into the host cell cytoplasm. The human parasite Plasmodium falciparum exports particularly large numbers of proteins, including proteins that establish a vesicular network allowing the trafficking of proteins onto the surface of irbcs that are responsible for tissue sequestration. Like P. falciparum, the rodent parasite P. berghei ANKA sequesters via irbc interactions with the host receptor CD36. We have applied proteomic, genomic, and reverse-genetic approaches to identify P. berghei proteins potentially involved in the transport of proteins to the irbc surface. A comparative proteomics analysis of P. berghei non-sequestering and sequestering parasites was used to determine changes in the irbc membrane associated with sequestration. Subsequent tagging experiments identified 13 proteins (Plasmodium export element (PEXEL)-positive as well as PEXEL-negative) that are exported into the irbc cytoplasm and have distinct localization patterns: a dispersed and/or patchy distribution, a punctate vesicle-like pattern in the cytoplasm, or a distinct location at the irbc membrane. Members of the PEXEL-negative BIR and PEXEL-positive Pb-fam-3 show a dispersed localization in the irbc cytoplasm, but not at the irbc surface. Two of the identified exported proteins are transported to the irbc membrane and were named erythrocyte membrane associated proteins. EMAP1 is a member of the PEXEL-negative Pb-fam-1 family, and EMAP2 is a PEXEL-positive protein encoded by a single copy gene; neither protein plays a direct role in sequestration. Our observations clearly indicate that P. berghei traffics a diverse range of proteins to different cellular locations via mechanisms that are analogous to those employed by P. falciparum. This information can be exploited to generate transgenic humanized rodent P. berghei parasites expressing chimeric P. berghei/P. falciparum proteins on the surface of rodent irbc, thereby opening new avenues for in vivo screening adjunct therapies that block sequestration.Malaria parasites invade and develop inside red blood cells, and extensive remodeling of the host cell is required in order for the parasite to take up nutrients and grow (1). In addition, infected red blood cells (irbcs)1 of several Plasmodium species adhere to endothelium lining blood capillaries, and this is achieved through modification of the irbc, specifically, alteration of the irbc membrane (2, 3). This active remodeling of the erythrocyte requires the export of parasite proteins into the host cell cytoplasm and their incorporation in the irbc membrane of the host cell (1, 2). Schizont-infected red blood cells of the rodent parasite P. berghei ANKA adhere to endothelial cells of the microvasculature, leading to the sequestration of irbcs in organs such as the lungs and adipose tissue (46). P. berghei irbcs adhere to the class II scavenger receptor CD36 (7), which is highly conserved in mammals and is the receptor most commonly used by irbcs infected with the human parasite P. falciparum (8). These observations suggest that P. berghei may export proteins onto the surface of irbcs in a fashion analogous to the processes employed by P. falciparum that expresses PfEMP1, the protein known to be responsible for P. falciparum irbc sequestration. However, P. berghei does not contain Pfemp1 orthologs or proteins with domains with clear homology to the domains of PfEMP1 (9), and the P. berghei proteins responsible for irbc cytoadherence and proteins involved in the transport of these proteins to the irbc membrane remain largely unknown. Recently we used a proteomic analysis of P. berghei ANKA irbc membranes to identify parasite proteins associated with the erythrocyte membrane, and we have demonstrated that the deletion of a single-copy gene of P. berghei that encodes a small exported protein known as SMAC results in strongly reduced irbc sequestration (6). No evidence was found for the presence of SMAC on the irbc surface, and therefore this protein is most likely involved in the transport or anchoring of other P. berghei proteins that directly interact with host receptors on endothelial cells.For P. falciparum, a large number of exported proteins have been predicted based on the presence of an N-terminal motif known as the Plasmodium export element (PEXEL) motif (10, 11). Many of these PEXEL-positive proteins belong to species-specific gene families. Comparison of PEXEL-positive proteins in different Plasmodium species suggested that P. falciparum expresses a significantly higher number of exported proteins than other Plasmodium species, which in part can be attributed to the expansion of P. falciparum–specific protein families, including those containing DnaJ or PHIST domains (1217). One explanation for the elevated number of exported proteins in P. falciparum is that they are necessary for export of the P. falciparum–specific protein PfEMP1 to the irbc surface (10). Comparisons of different Plasmodium exportomes have mainly focused on identifying orthologs of the PEXEL-positive proteins of P. falciparum in the other species (14, 15, 18). For example, of the >500 PEXEL-positive P. falciparum proteins, only between 11 and 33 had orthologs in P. berghei (14, 15, 19). However, such an approach might underestimate the total number of exported proteins. A recent hidden Markov model (HMM) analysis of the PEXEL motif for P. berghei proteins identified at least 75 PEXEL-positive P. berghei proteins (6). Moreover, in different Plasmodium species, a number of exported proteins have been described that are PEXEL-negative, indicating that alternative export pathways might exist that are independent of the presence of a PEXEL motif (20, 21). It has been suggested that in species with a small number of PEXEL-positive proteins, PEXEL-negative exported proteins play a more prominent role in host cell remodeling (21). An example of a PEXEL-negative exported protein family is the large PIR family of proteins, which are expressed by rodent Plasmodium species (9, 22), the monkey parasite P. knowlesi (23), and the human parasite P. vivax (24, 25).To date, export to the irbc cytosol has been shown for only a few P. berghei proteins (i.e. several members of the BIR family; TIGR01590) (6), two members of the ETRAMP family (26), and two proteins encoded by a single copy gene, SMAC and IBIS1 (6, 27). In this study, comparative proteomic, genomic, and reverse-genetic approaches have been used to identify novel exported proteins of P. berghei. We report proteome analyses of samples enriched for proteins associated with membranes of irbcs from both sequestering P. berghei ANKA and non-sequestering P. berghei K173 parasites, and we also present analyses of the full genome sequence of a non-sequestering P. berghei K173 line. Fluorescent tagging of parasite proteins selected from the proteome and genome analyses identified a number of novel P. berghei ANKA proteins that are exported into the irbc cytoplasm. We report for the first time the export of members of the PEXEL-negative Pb-fam-1 gene family (pyst-a; TIGR01599) and show that two proteins are transported to the P. berghei ANKA irbc membrane. This is the first comprehensive study of exported proteins of P. berghei that has been validated via the generation of a large number of transgenic P. berghei ANKA parasites expressing tagged proteins and has shown the export of both PEXEL-positive and PEXEL-negative proteins to the irbc cytoplasm. The identification of P. berghei ANKA proteins exported to the irbc membrane and proteins involved in sequestration suggests the possibility of developing “humanized” small animal models for the in vivo analysis of the sequestration properties of P. falciparum proteins that would express (domains of) P. falciparum proteins on the surface of rodent irbcs (4, 6).  相似文献   

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

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