首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 46 毫秒
1.
Ornithine decarboxylase (ODC) is the first enzyme involved in polyamine biosynthesis, and it catalyzes the decarboxylation of ornithine to putrescine. ODC is a dimeric enzyme, whereas antizyme inhibitor (AZI), a positive regulator of ODC that is homologous to ODC, exists predominantly as a monomer and lacks decarboxylase activity. The goal of this paper was to identify the essential amino acid residues that determine the dimerization of AZI. The nonconserved amino acid residues in the putative dimer interface of AZI (Ser-277, Ser-331, Glu-332, and Asp-389) were substituted with the corresponding residues in the putative dimer interface of ODC (Arg-277, Tyr-331, Asp-332, and Tyr-389, respectively). Analytical ultracentrifugation analysis was used to determine the size distribution of these AZI mutants. The size-distribution analysis data suggest that residue 331 may play a major role in the dimerization of AZI. Mutating Ser-331 to Tyr in AZI (AZI-S331Y) caused a shift from a monomer configuration to a dimer. Furthermore, in comparison with the single mutant AZI-S331Y, the AZI-S331Y/D389Y double mutant displayed a further reduction in the monomer-dimer Kd, suggesting that residue 389 is also crucial for AZI dimerization. Analysis of the triple mutant AZI-S331Y/D389Y/S277R showed that it formed a stable dimer (Kd value = 1.3 μm). Finally, a quadruple mutant, S331Y/D389Y/S277R/E332D, behaved as a dimer with a Kd value of ∼0.1 μm, which is very close to that of the human ODC enzyme. The quadruple mutant, although forming a dimer, could still be disrupted by antizyme (AZ), further forming a heterodimer, and it could rescue the AZ-inhibited ODC activity, suggesting that the AZ-binding ability of the AZI dimer was retained.Polyamines (putrescine, spermidine, and spermine) have been shown to have both structural and regulatory roles in protein and nucleic acid biosynthesis and function (13). Ornithine decarboxylase (ODC,3 EC 4.1.1.17) is a central regulator of cellular polyamine synthesis (reviewed in Refs. 1, 4, 5). This enzyme catalyzes the pyridoxal 5-phosphate (PLP)-dependent decarboxylation of ornithine to putrescine, and it is the first and rate-limiting enzyme in polyamine biosynthesis (2, 3, 6, 7). ODC and polyamines play important roles in a number of biological functions, including embryonic development, cell cycle, proliferation, differentiation, and apoptosis (815). They also have been associated with human diseases and a variety of cancers (1626). Because the regulation of ODC and polyamine content is critical to cell proliferation (11), as well as in the origin and progression of neoplastic diseases (23, 24), ODC has been identified as an oncogenic enzyme, and the inhibitors of ODC and the polyamine pathway are important targets for therapeutic intervention in many cancers (6, 11).ODC is ubiquitously found in organisms ranging from bacteria to humans. It contains 461 amino acid residues in each monomer and is a 106-kDa homodimer with molecular 2-fold symmetry (27, 28). Importantly, ODC activity requires the formation of a dimer (2931). X-ray structures of the ODC enzyme reveal that this dimer contains two active sites, both of which are formed at the interface between the N-terminal domain of one monomer, which provides residues involved in PLP interactions, and the C-terminal domain of the other subunit, which provides the residues that interact with substrate (27, 3241).ODC undergoes a unique ubiquitin-independent proteasomal degradation via a direct interaction with the regulatory protein antizyme (AZ). Binding of AZ promotes the dissociation of the ODC homodimers and targets ODC for degradation by the 26 S proteasome (4246). Current models of antizyme function indicate that increased polyamine levels promote the fidelity of the AZ mRNA translational frameshift, leading to increased concentrations of AZ (47). The AZ monomer selectively binds to dimeric ODC, thereby inactivating ODC by forming inactive AZ-ODC heterodimers (44, 4850). AZ acts as a regulator of polyamine metabolism that inhibits ODC activity and polyamine transport, thus restricting polyamine levels (4, 5, 51, 52). When antizymes are overexpressed, they inhibit ODC and promote ubiquitin-independent proteolytic degradation of ODC. Because elevated ODC activity is associated with most forms of human malignancies (1), it has been suggested that antizymes may function as tumor suppressors.In contrast to the extensive studies on the oncogene ODC, the endogenous antizyme inhibitor (AZI) is less well understood. AZI is homologous to the enzyme ODC. It is a 448-amino acid protein with a molecular mass of 50 kDa. However, despite the homology between these proteins, AZI does not possess any decarboxylase activity. It binds to antizyme more tightly than does ODC and releases ODC from the ODC-antizyme complex (53, 54). Both the AZI and AZ proteins display rapid ubiquitin-dependent turnover within a few minutes to 1 h in vivo (5). However, AZ binding actually stabilizes AZI by inhibiting its ubiquitination (55).AZI, which inactivates all members of the AZ family (53, 56), restores ODC activity (54), and prevents the proteolytic degradation of ODC, may play a role in tumor progression. It has been reported that down-regulation of AZI is associated with the inhibition of cell proliferation and reduced ODC activity, presumably through the modulation of AZ function (57). Moreover, overexpression of AZI has been shown to increase cell proliferation and promote cell transformation (5860). Furthermore, AZI is capable of direct interaction with cyclin D1, preventing its degradation, and this effect is at least partially independent of AZ function (60, 61). These results demonstrate a role for AZI in the positive regulation of cell proliferation and tumorigenesis.It is now known that ODC exists as a dimer and that AZI may exist as a monomer physiologically (62). Fig. 1 shows the dimeric structures of ODC (Fig. 1A) and AZI (Fig. 1B). Although structural studies indicate that both ODC and AZI crystallize as dimers, the dimeric AZI structure has fewer interactions at the dimer interface, a smaller buried surface area, and a lack of symmetry of the interactions between residues from the two monomers, suggesting that the AZI dimer may be nonphysiological (62). In this study, we identify the critical amino acid residues governing the difference in dimer formation between ODC and AZI. Our preliminary studies using analytical ultracentrifugation indicated that ODC exists as a dimer, whereas AZI exists in a concentration-dependent monomer-dimer equilibrium. Multiple sequence alignments of ODC and AZI from various species have shown that residues 277, 331, 332, and 389 are not conserved between ODC and AZI (Open in a separate windowFIGURE 1.Crystal structure and the amino acid residues at the dimer interface of human ornithine decarboxylase (hODC) and mouse antizyme inhibitor (mAZI). A, homodimeric structure of human ODC with the cofactor PLP analog, LLP (Protein Data Bank code 1D7K). B, putative dimeric structure of mouse AZI (Protein Data Bank code 3BTN). The amino acid residues in the dimer interface are shown as a ball-and-stick model. The putative AZ-binding site is colored in cyan. This figure was generated using PyMOL (DeLano Scientific LLC, San Carlos, CA).

TABLE 1

Amino acid residues at the dimer interface of human ODC and AZI
Human ODCResidueHuman AZI
Nonconserved
    Arg277Ser
    Tyr331Ser
    Asp332Glu
    Tyr389Asp

Conserved
    Asp134Asp
    Lys169Lys
    Lys294Lys
    Tyr323Tyr
    Asp364Asp
    Gly387Gly
    Phe397Phe
Open in a separate window  相似文献   

2.
3.
Over the last years, large scale proteomics studies have generated a wealth of information of biomolecular complexes. Adding the structural dimension to the resulting interactomes represents a major challenge that classical structural experimental methods alone will have difficulties to confront. To meet this challenge, complementary modeling techniques such as docking are thus needed. Among the current docking methods, HADDOCK (High Ambiguity-Driven DOCKing) distinguishes itself from others by the use of experimental and/or bioinformatics data to drive the modeling process and has shown a strong performance in the critical assessment of prediction of interactions (CAPRI), a blind experiment for the prediction of interactions. Although most docking programs are limited to binary complexes, HADDOCK can deal with multiple molecules (up to six), a capability that will be required to build large macromolecular assemblies. We present here a novel web interface of HADDOCK that allows the user to dock up to six biomolecules simultaneously. This interface allows the inclusion of a large variety of both experimental and/or bioinformatics data and supports several types of cyclic and dihedral symmetries in the docking of multibody assemblies. The server was tested on a benchmark of six cases, containing five symmetric homo-oligomeric protein complexes and one symmetric protein-DNA complex. Our results reveal that, in the presence of either bioinformatics and/or experimental data, HADDOCK shows an excellent performance: in all cases, HADDOCK was able to generate good to high quality solutions and ranked them at the top, demonstrating its ability to model symmetric multicomponent assemblies. Docking methods can thus play an important role in adding the structural dimension to interactomes. However, although the current docking methodologies were successful for a vast range of cases, considering the variety and complexity of macromolecular assemblies, inclusion of some kind of experimental information (e.g. from mass spectrometry, nuclear magnetic resonance, cryoelectron microscopy, etc.) will remain highly desirable to obtain reliable results.Proteins are the wheels and millstones of the complex machinery that underlies human life. Catalyzing a huge diversity of chemical processes, proteins work in close association with other biomolecules: nucleic acids, sugars, lipids, and other proteins. This huge network of protein interactions enables the cell to respond quickly to changes in the environment, such as temperature, oxygen, or nutrient concentration. However, to fully understand this network, insights at the atomic level are needed.In the wake of the elucidation of the human genome (1, 2), many structural genomics projects are solving the structures of what is now becoming a considerable fraction of the human proteome (3). These projects are now moving to the next level, which is solving the atomic resolution structures of protein complexes. However, this is a challenge that is considerably greater than obtaining the structures of single proteins. First of all, a protein can take part in 10 interactions on average; thus, the number of complexes is expected to be at least an order of magnitude larger than the proteome, and their composition can even vary over time. Second, associations between subunits in protein complexes are often weak and reversible, which make purification and crystallization difficult. Finally, there are some very well studied classes of interactions, such as enzyme-inhibitor, antibody-antigen, and GTPase-GAP (GTPase-activating protein) interactions, but these classes represent binary interactions between proteins. In contrast, many of the most important functions in the cell are carried out by large, dynamic molecular assemblies, such as the ribosome, the proteasome, the spliceosome, RNA polymerases, and the nuclear pore complex (4, 5). For such assemblies, high resolution methods such as x-ray crystallography and NMR spectroscopy often provide atomic level information at the level of individual subunits or subcomplexes, but they typically encounter difficulties at the level of the full complex.Fortunately, low resolution information about protein complexes can often be obtained. Affinity purification (6, 7) followed by mass spectrometry is a high throughput technique to study the composition of a complex. However, dissociation inside the mass spectrometer can be a problem for transient or unstable complexes in which case chemical cross-linking can help. Once the composition of the complex is known, there is a variety of experimental techniques available to obtain structural information on the complex. The most detailed information can be gathered by using data obtained from various NMR experiments, for example chemical shift perturbations (8) or residual dipolar couplings (9); unfortunately, NMR is limited to complexes that are fairly small in size, making its applicability in the context of large assemblies less suited. Techniques that provide information about the shape of a protein complex, such as small angle x-ray scattering (SAXS),1 cryoelectron tomography, and single molecule cryoelectron microscopy (cryo-EM), are more suited to characterize large complexes. Unfortunately, all of these techniques suffer from limitations in resolution that are either fundamental or caused by structural heterogeneities of the complex.A well known approach to obtain information on residues at an interface is site-directed mutagenesis (10). In principle, a loss of binding affinity indicates that the mutated residue mediates the interaction, although the reverse is not true. Also, one must take care of secondary effects, such as unfolding or conformational change caused by the mutation. Apart from that, very detailed information about interface residues can be obtained by extensive mutagenesis experiments, such as alanine scanning and double mutant cycles. Mass spectrometry offers the opportunity to get peptide level or residue level information about protein interfaces by accurate mass measurements of peptides from the protein complex, generated either a priori through proteolytic cleavage, or inside the mass spectrometer (MS/MS). For example, interface residues can be identified as residues that undergo slower hydrogen/deuterium exchange upon complex formation. This process can be monitored at the peptide level by mass spectrometry (or in smaller complexes, at the residue level by NMR), although this method is very sensitive to noise caused by conformational changes upon binding. In the same way, radical probe MS (RP-MS) uses differences in oxidation of residues by hydroxyl radicals generated in the mass spectrometer to identify interface residues. Finally, chemical cross-linking followed by MS can provide direct information about residue contact sites between different binding partners of the complex. Several cross-linking reagents can provide complementary information. However, it has been reported that the cross-linkers may disrupt the structure of the protein complex and that care should therefore be taken to interpret the results (11).There is a need for computational approaches to translate this low resolution information into atomic resolution models that can provide functional and mechanistic insights. One of the most promising approaches is docking, the prediction of the structure of a complex starting from the free, unbound structures of its constituents. In recent years, docking methods have made much progress in the blind prediction of the structure of protein complexes as seen in the recent rounds of the critical assessment of prediction of interactions (CAPRI) experiment (12, 13). Most docking methods are ab initio, which means that experimental data are not required. However, it is possible in several ab initio methods to use experimentally determined interface residues in the docking: in MolFit (14, 15) and ATTRACT (16, 17), it is possible to up-weight the interaction scores of interface residues; in ZDOCK (18, 19), it is possible to block non-interface residues; and in PatchDock (20, 21), ZDOCK, pyDock (22, 23), and several other methods, it is possible to filter the docking results based on experimental information. Next to purely ab initio approaches, there are also methods that make use of different types experimental information, for example PROXIMO (24), based on RP-MS data, and MultiFit (25), a hybrid fitting/docking approach based on electron microscopy data.A method that distinguishes itself from the variety of above mentioned docking approaches is HADDOCK (2628). In HADDOCK, the docking can be driven by a variety of experimental data using information about interface, contacts, and relative orientations inside a complex simultaneously. Originally developed for NMR data, HADDOCK is able to deal with a large variety of experimental data as shown in 2628) and “Materials and Methods” for more details.) HADDOCK has performed very well in translating these data into structures and structural models. More than 60 Protein Data Bank structures calculated using HADDOCK have been deposited to date as experimental structures in the Protein Data Bank (29). Moreover, HADDOCK has shown a strong performance in CAPRI. Finally, HADDOCK is a general purpose program that can integrate many kinds of data, but even with a single source of data it is able to perform as well as more specialized programs: for example, HADDOCK was able to closely reproduce the NMR-calculated E2A-HPr complex using only chemical shift perturbation data. For the ribonuclease S-protein-peptide complex (Protein Data Bank code 1J80 (30)) for which RP-MS data are available, PROXIMO was able to closely reproduce the crystal structure (root mean square deviation (r.m.s.d.) of the top scoring model from the reference crystal structure is 1.26 Å); using the same data, HADDOCK could get even closer with an r.m.s.d. of only 0.68 Å from the crystal structure (results not shown).

Table I

Various experimental data that can be incorporated into HADDOCK
Experimental dataHADDOCK representation
Mutagenesis dataActive and passive residues
Hydrogen/deuterium exchange dataActive and passive residues
Bioinformatics interface predictionsActive and passive residues
Mass spectrometry data
    Cross-linking dataCustom CNS restraints
    Radical probe mass spectrometryActive and passive residues
    Limited proteolysis mass spectrometryActive and passive residues or directly as an MTMDAT-generated HADDOCK parameter file
NMR data
    Chemical shift perturbation dataActive and passive residues
    Cross-saturation experimentsActive and passive residues
    Residual dipolar couplingsDirectly
    Diffusion anisotropy restraintsDirectly
    NOEsa as custom CNS restraintsCustom CNS restraints
    Dihedral anglesDirectly
    Hydrogen bondsDirectly
    Paramagnetic restraintsUnder development
Shape data
    SAXSUnder development
    EMUnder development
Open in a separate windowa Nuclear Overhauser effects.Most docking methods are designed to deal with just two molecules, making their application limited with regard to large macromolecular assemblies. In most programs, multicomponent complexes can be assembled by adding each component one at a time, whereas simultaneous docking of the whole complex is typically not possible. Recently five ab initio docking programs (MolFit (31, 32), ClusPro (33), Rosetta (34), M-ZDOCK (35), and SymmDock (36)) gave birth to specific versions for the prediction of the symmetric multimers. Among these programs, MolFit, ClusPro, and Rosetta perform a rotational/translational search about the proper symmetry axes. These programs can deal with different types of cyclic and dihedral symmetries. Different than the other two, Rosetta is able to assemble complexes having helical and icosahedral symmetries. M-ZDOCK and SymmDock are suited for the prediction of macromolecules with cyclic symmetries. However, the ability to deal with arbitrary large molecular assemblies is currently rare. CombDock (37), which was developed by the team of SymmDock, can build hetero-oligomer complexes, but it does not have a symmetry option. Only HADDOCK can deal with molecular complexes that are hetero-oligomers or homo-oligomers with arbitrary symmetry operators between and within each component.The flexibility of HADDOCK comes at a price: it requires the user to have the structure calculation program CNS (38) installed and a considerable degree of expertise in its usage and molecular modeling in general, and it requires a cluster of computers. To alleviate this problem and to open up HADDOCK for a wide community, we have recently developed the HADDOCK web server (27). The server offers multiple web interfaces, ranging from very simple and user-friendly to very powerful and flexible, exposing the full range of HADDOCK options to the expert user. However, up until now, the HADDOCK server was unable to deal with more than two molecules. Here we present a novel web interface for multibody docking of complexes. Like the HADDOCK program itself, the server supports the docking of up to six molecules simultaneously; all HADDOCK options, including symmetry restraints, are made available to the user. Even larger assemblies can in principle be modeled if the docking is performed in an incremental way. Here we demonstrate the performance of the multibody server on a small benchmark comprising complexes of various symmetries and increasing numbers of components (from three to five). To drive the docking, bioinformatics interface predictions and/or available experimental information were used. The HADDOCK server is available on line. http://haddock.chem.uu.nl.  相似文献   

4.
5.
Nitrogenase-like light-independent protochlorophyllide oxidoreductase (DPOR) is involved in chlorophyll biosynthesis. Bacteriochlorophyll formation additionally requires the structurally related chlorophyllide oxidoreductase (COR). During catalysis, homodimeric subunit BchL2 or ChlL2 of DPOR transfers electrons to the corresponding heterotetrameric catalytic subunit, (BchNB)2 or (ChlNB)2. Analogously, subunit BchX2 of the COR enzymes delivers electrons to subunit (BchYZ)2. Various chimeric DPOR enzymes formed between recombinant subunits (BchNB)2 and BchL2 from Chlorobaculum tepidum or (ChlNB)2 and ChlL2 from Prochlorococcus marinus and Thermosynechococcus elongatus were found to be enzymatically active, indicating a conserved docking surface for the interaction of both DPOR protein subunits. Biotin label transfer experiments revealed the interaction of P. marinus ChlL2 with both subunits, ChlN and ChlB, of the (ChlNB)2 tetramer. Based on these findings and on structural information from the homologous nitrogenase system, a site-directed mutagenesis approach yielded 10 DPOR mutants for the characterization of amino acid residues involved in protein-protein interaction. Surface-exposed residues Tyr127 of subunit ChlL, Leu70 and Val107 of subunit ChlN, and Gly66 of subunit ChlB were found essential for P. marinus DPOR activity. Next, the BchL2 or ChlL2 part of DPOR was exchanged with electron-transferring BchX2 subunits of COR and NifH2 of nitrogenase. Active chimeric DPOR was generated via a combination of BchX2 from C. tepidum or Roseobacter denitrificans with (BchNB)2 from C. tepidum. No DPOR activity was observed for the chimeric enzyme consisting of NifH2 from Azotobacter vinelandii in combination with (BchNB)2 from C. tepidum or (ChlNB)2 from P. marinus and T. elongatus, respectively.Chlorophyll and bacteriochlorophyll biosynthesis, as well as nitrogen fixation, are essential biochemical processes developed early in the evolution of life (1). During biological fixation of nitrogen, nitrogenase catalyzes the reduction of atmospheric dinitrogen to ammonia (2). Enzyme systems homologous to nitrogenase play a crucial role in the formation of the chlorin and bacteriochlorin ring system of chlorophylls (Chl)2 and bacteriochlorophylls (Bchl) (3, 4) (Fig. 1a). For the synthesis of both Chl and Bchl, the stereospecific reduction of the C-17-C-18 double bond of ring D of protochlorophyllide (Pchlide) catalyzed by the nitrogenase-like enzyme light-independent (dark-operative) protochlorophyllide oxidoreductase (DPOR) results in the formation of chlorophyllide (Chlide) (Fig. 1a, left) (5, 6). DPOR enzymes consist of three protein subunits which are designated BchN, BchB and BchL in Bchl-synthesizing organisms and ChlN, ChlB and ChlL in Chl-synthesizing organisms. A second reduction step at ring B (C-7-C-8) unique to the synthesis of Bchl converts the chlorin Chlide into a bacteriochlorin ring structure to form bacteriochlorophyllide (Bchlide) (Fig. 1a, right, Bchlide). This reaction is catalyzed by another nitrogenase-like enzyme, termed chlorophyllide oxidoreductase (COR) (7). COR enzymes are composed of subunits BchY, BchZ, and BchX.Open in a separate windowFIGURE 1.Comparison of the three subunit enzymes DPOR, COR, and nitrogenase. a, during Chl and Bchl biosynthesis, ring D is stereospecifically reduced by the nitrogenase-like enzyme DPOR (subunit composition BchL2/(BchNB)2 or ChlL2/(ChlNB)2) leading to the chlorin Chlide. Subunits N, B, and L are named ChlN, ChlB, and ChlL in Chl-synthesizing organisms and BchN, BchB, and BchL in Bchl-synthesizing organisms. The synthesis of Bchl additionally requires the stereospecific B ring reduction by a second nitrogenase-like enzyme called COR, with the subunit composition BchX2/(BchYZ)2. COR catalyzes the formation of the bacteriochlorin Bchlide. Subunits Y, Z, and X of the COR enzyme are named BchY, BchZ, and BchX. b, the homologous nitrogenase complex has the subunit composition NifH2/(NifD/NifK)2. Rings A–E and the carbon atoms are designated according to IUPAC nomenclature (41). R is either a vinyl or an ethyl moiety. The position marked by an asterisk indicates either a vinyl or a hydroxyethyl moiety (42).All subunits share significant amino acid sequence homology to the corresponding subunits of nitrogenase, which are designated NifD, NifK, and NifH, respectively (1) (compare Fig. 1, a and b). Whereas subunits BchL or ChlL, BchX and NifH exhibit a sequence identity at the amino acid level of ∼33%, subunits BchN or ChlN, BchY, NifD, and BchB or ChlB, BchZ, and NifK, respectively, show lower sequence identities of ∼15% (1). For all enzymes a common oligomeric protein architecture has been proposed consisting of the heterotetrameric complexes (BchNB)2 or (ChlNB)2, (BchYZ)2, and (NifD/NifK)2, which are completed by a homodimeric protein subunit BchL2 or ChlL2, BchX2, and NifH2, respectively (compare Fig. 1, a and b) (3, 7, 8).Nitrogenase is a well characterized protein complex that catalyzes the reduction of nitrogen to ammonia in a reaction that requires at least 16 molecules of MgATP (2, 9, 10). During nitrogenase catalysis, subunit NifH2 (Fe protein) associates with and dissociates from the (NifD/NifK)2 complex (MoFe protein). Binding, hydrolysis of MgATP and structural rearrangements are coupled to sequential intersubunit electron transfer. For this purpose, NifH2 contains an ATP-binding motif and an intersubunit [4Fe-4S] cluster coordinated by two cysteine residues from each NifH monomer (1, 11). Electrons from this [4Fe-4S] cluster are transferred via a [8Fe-7S] cluster (P-cluster) onto the [1Mo-7Fe-9S-X-homocitrate] cluster (MoFe cofactor). Both of the latter clusters are located on (NifD/NifK)2, where dinitrogen is reduced to ammonia (10). Three-dimensional structures of NifH2 in complex with (NifD/NifK)2 revealed a detailed picture of the dynamic interaction of both subcomplexes (8, 12).Based on biochemical and bioinformatic approaches, it has been proposed that the initial steps of DPOR reaction strongly resemble nitrogenase catalysis. Key amino acid residues essential for DPOR function have been identified by mutagenesis of the enzyme from Chlorobaculum tepidum (formerly denoted as Chlorobium tepidum) (3). The catalytic mechanism of DPOR includes the electron transfer from a “plant-type” [2Fe-2S] ferredoxin onto the dimeric DPOR subunit, BchL2, carrying an intersubunit [4Fe-4S] redox center coordinated by Cys97 and Cys131 in C. tepidum. Analogous to nitrogenase, Lys10 in the phosphate-binding loop (P-loop) and Leu126 in the switch II region of DPOR were found essential for DPOR catalysis. Moreover, it was shown that the BchL2 protein from C. tepidum does not form a stable complex with the catalytic (BchNB)2 subcomplex. Therefore, a transient interaction responsible for the electron transfer onto protein subunit (BchNB)2 has been proposed (3).The subsequent [Fe-S] cluster-dependent catalysis and the specific substrate recognition at the active site located on subunit (BchNB)2 are unrelated to nitrogenase. The (BchNB)2 subcomplex was shown to carry a second [4Fe-4S] cluster, which was proposed to be ligated by Cys21, Cys46, and Cys103 of the BchN subunit and Cys94 of subunit BchB (C. tepidum numbering) (3). No evidence for any type of additional cofactor was obtained from biochemical and EPR spectroscopic analyses (5, 13). Thus, despite the same common oligomeric architecture, the catalytic subunits (BchNB)2 and (ChlNB)2 clearly differ from the corresponding nitrogenase complex, as no molybdenum-containing cofactor or P-cluster equivalent is employed (5, 14). From these results it was concluded that electrons from the [4Fe-4S] cluster of (BchNB)2 or (ChlNB)2 are transferred directly onto the Pchlide substrate at the active site of DPOR.The second nitrogenase-like enzyme, COR, catalyzes the reduction of ring B of Chlide during the biosynthesis of Bchl (7). Therefore, an accurate discrimination of the ring systems of the individual substrates is required. COR subunits share an overall amino acid sequence identity of 15–22% for BchY and BchZ and 31–35% for subunit BchX when compared with the corresponding DPOR subunits (supplemental Figures S2–S4). In amino acid sequence alignments of BchX proteins with the closely related BchL or ChlL subunits of DPOR, both cysteinyl ligands responsible for [4Fe-4S] cluster formation and residues for ATP binding are conserved (1). Furthermore, all cysteinyl residues characterized as ligands for a catalytic [4Fe-4S] cluster in (BchNB)2 or (ChlNB)2 are conserved in the sequences of subunits BchY and BchZ of COR (7). These findings correspond to a recent EPR study in which a characteristic signal for a [4Fe-4S] cluster was obtained for the COR subunit BchX2 as well as for subunit (BchYZ)2 (15). These results indicate that the catalytic mechanism of COR strongly resembles DPOR catalysis. In vitro assays for nitrogenase as well as for DPOR and COR make use of the artificial electron donor dithionite in the presence of high concentrations of ATP (7, 16, 17).

TABLE 1

Amino acid sequence identities of the individual subunits of DPOR, COR, and nitrogenaseAmino acid sequences of the individual subunits of DPOR, COR, and nitrogenase employed in the present study (compareFig. 3A) were aligned by using the ClustalW method in MegAlign (DNASTAR), and sequence identities were calculated.
DPOR
COR
Nitrogenase
NBLYZXNifDNifKNifH
DPOR
    N37–5815–1812–20
    B34–6215–2214–18
    L51–6931–3531–38

COR
    Y35–7813–15
    Z39–8111–16
    X42–8329–36

Nitrogenase
    NifD17–70
    NifK37–58
    NifH67–75
Open in a separate windowIn this study, we investigated the transient interaction of the dimeric subunit BchL2 or ChlL2 with the heterotetrameric (BchNB)2 or (ChlNB)2 complex, which is essential for DPOR catalysis. We make use of the individually purified DPOR subunits BchL2 and (BchNB)2 from the green sulfur bacterium C. tepidum and ChlL2 and (ChlNB)2 from the prochlorophyte Prochlorococcus marinus and from the cyanobacterium Thermosynechococcus elongatus. The individual combination of (BchNB)2 or (ChlNB)2 complexes and BchL2 or ChlL2 proteins from these organisms resulted in catalytically active chimeras of DPOR. These results enabled us to propose conserved regions of the postulated docking surface, which were subsequently verified in a mutagenesis study. To elucidate the potential evolution of the electron-transferring subunit of nitrogenase and nitrogenase-like enzymes, we also analyzed chimeric enzymes consisting of DPOR subunits (BchNB)2 or (ChlNB)2 in combination with subunits BchX2 from C. tepidum and R. denitrificans of the COR enzyme and with subunit NifH2 of nitrogenase from Azotobacter vinelandii, respectively.  相似文献   

6.
Forty-five different point mutations in POLG, the gene encoding the catalytic subunit of the human mitochondrial DNA polymerase (pol γ), cause the early onset mitochondrial DNA depletion disorder, Alpers syndrome. Sequence analysis of the C-terminal polymerase region of pol γ revealed a cluster of four Alpers mutations at highly conserved residues in the thumb subdomain (G848S, c.2542g→a; T851A, c.2551a→g; R852C, c.2554c→t; R853Q, c.2558g→a) and two Alpers mutations at less conserved positions in the adjacent palm subdomain (Q879H, c.2637g→t and T885S, c.2653a→t). Biochemical characterization of purified, recombinant forms of pol γ revealed that Alpers mutations in the thumb subdomain reduced polymerase activity more than 99% relative to the wild-type enzyme, whereas the palm subdomain mutations retained 50–70% wild-type polymerase activity. All six mutant enzymes retained physical and functional interaction with the pol γ accessory subunit (p55), and none of the six mutants exhibited defects in misinsertion fidelity in vitro. However, differential DNA binding by these mutants suggests a possible orientation of the DNA with respect to the polymerase during catalysis. To our knowledge this study represents the first structure-function analysis of the thumb subdomain in pol γ and examines the consequences of mitochondrial disease mutations in this region.As the only DNA polymerase found in animal cell mitochondria, DNA polymerase γ (pol γ)3 bears sole responsibility for DNA synthesis in all replication and repair transactions involving mitochondrial DNA (1, 2). Mammalian cell pol γ is a heterotrimeric complex composed of one catalytic subunit of 140 kDa (p140) and two 55-kDa accessory subunits (p55) that form a dimer (3). The catalytic subunit contains an N-terminal exonuclease domain connected by a linker region to a C-terminal polymerase domain. Whereas the exonuclease domain contains essential motifs I, II, and III for its activity, the polymerase domain comprising the thumb, palm, and finger subdomains contains motifs A, B, and C that are crucial for polymerase activity. The catalytic subunit is a family A DNA polymerase that includes bacterial pol I and T7 DNA polymerase and possesses DNA polymerase, 3′ → 5′ exonuclease, and 5′-deoxyribose phosphate lyase activities (for review, see Refs. 1 and 2). The 55-kDa accessory subunit (p55) confers processive DNA synthesis and tight binding of the pol γ complex to DNA (4, 5).Depletion of mtDNA as well as the accumulation of deletions and point mutations in mtDNA have been observed in several mitochondrial disorders (for review, see Ref. 6). mtDNA depletion syndromes are caused by defects in nuclear genes responsible for replication and maintenance of the mitochondrial genome (7). Mutation of POLG, the gene encoding the catalytic subunit of pol γ, is frequently involved in disorders linked to mutagenesis of mtDNA (8, 9). Presently, more than 150 point mutations in POLG are linked with a wide variety of mitochondrial diseases, including the autosomal dominant (ad) and recessive forms of progressive external ophthalmoplegia (PEO), Alpers syndrome, parkinsonism, ataxia-neuropathy syndromes, and male infertility (tools.niehs.nih.gov/polg) (9).Alpers syndrome, a hepatocerebral mtDNA depletion disorder, and myocerebrohepatopathy are rare heritable autosomal recessive diseases primarily affecting young children (1012). These diseases generally manifest during the first few weeks to years of life, and symptoms gradually develop in a stepwise manner eventually leading to death. Alpers syndrome is characterized by refractory seizures, psychomotor regression, and hepatic failure (11, 12). Mutation of POLG was first linked to Alpers syndrome in 2004 (13), and to date 45 different point mutations in POLG (18 localized to the polymerase domain) are associated with Alpers syndrome (9, 14, 15). However, only two Alpers mutations (A467T and W748S, both in the linker region) have been biochemically characterized (16, 17).During the initial cloning and sequencing of the human, Drosophila, and chicken pol γ genes, we noted a highly conserved region N-terminal to motif A in the polymerase domain that was specific to pol γ (18). This region corresponds to part of the thumb subdomain that tracks DNA into the active site of both Escherichia coli pol I and T7 DNA polymerase (1921). A high concentration of disease mutations, many associated with Alpers syndrome, is found in the thumb subdomain.Here we investigated six mitochondrial disease mutations clustered in the N-terminal portion of the polymerase domain of the enzyme (Fig. 1A). Four mutations (G848S, c.2542g→a; T851A, c.2551a→g; R852C, c.2554c→t; R853Q, c.2558g→a) reside in the thumb subdomain and two (Q879H, c.2637g→t and T885S, c.2653a→t) are located in the palm subdomain. These mutations are associated with Alpers, PEO, mitochondrial myopathy, encephalopathy, lactic acidosis, and stroke-like episodes (MELAS), ataxia-neuropathy syndrome, Leigh syndrome, and myocerebrohepatopathy (
POLG mutationDiseaseGeneticsReference
G848SAlpers syndromeIn trans with A467T, Q497H, T251I-P587L, or W748S-E1143G in Alpers syndrome15, 35, 4350
Leigh syndromeIn trans with R232H in Leigh syndrome49
MELASIn trans with R627Q in MELAS38
PEO with ataxia-neuropathyIn trans with G746S and E1143G in PEO with ataxia50
PEOIn trans with T251I and P587L in PEO51, 52
T851AAlpers syndromeIn trans with R1047W48, 53
In trans with H277C
R852CAlpers syndromeIn trans with A467T14, 48, 50
In cis with G11D and in trans with W748S-E1143G or A467T
Ataxia-neuropathyIn trans with G11D-R627Q15
R853QMyocerebrohepatopathyIn trans with T251I-P587L15
Q879HAlpers syndrome with valproate-induced hepatic failureIn cis with E1143G and in trans with A467T-T885S35, 54
T885SAlpers syndrome with valproate-induced hepatic failureIn cis with A467T and in trans with Q879H-E1143G35, 54
Open in a separate windowOpen in a separate windowFIGURE 1.POLG mutations characterized in this study. A, the location of the six mutations characterized is shown in red in the primary sequence of pol γ. Four mutations, the G848S, T851A, R852C, and R853Q, are located in the thumb domain, whereas two mutations, the Q879H and T885S, are in the palm domain of the polymerase region. B, sequence alignment of pol γ from yeast to humans. The amino acids characterized in this study are shown in red. Yellow-highlighted amino acids are highly conserved, and blue-highlighted amino acids are moderately conserved.  相似文献   

7.
Functional Characterization of Naturally Occurring Variants of Human Hepatitis B Virus Containing the Core Internal Deletion Mutation     
Thomas Ta-Tung Yuan  Min-Hui Lin  Sui Min Qiu  Chiaho Shih 《Journal of virology》1998,72(3):2168-2176
  相似文献   

8.
Single-cell-type Proteomics: Toward a Holistic Understanding of Plant Function     
Shaojun Dai  Sixue Chen 《Molecular & cellular proteomics : MCP》2012,11(12):1622-1630
  相似文献   

9.
Ca2+/Calmodulin-dependent Protein Kinase IV Links Group I Metabotropic Glutamate Receptors to Fragile X Mental Retardation Protein in Cingulate Cortex     
Hansen Wang  Hotaka Fukushima  Satoshi Kida    Min Zhuo 《The Journal of biological chemistry》2009,284(28):18953-18962
  相似文献   

10.
Repression of recA Induction by RecX Is Independent of the RecA Protein in Deinococcus radiodurans     
Duohong Sheng  Mingfeng Li  Jiandong Jiao  Xiehuang Sheng  Wenqiang Deng  Yuejin Hua 《Journal of bacteriology》2010,192(13):3540-3544
  相似文献   

11.
Biological Activity of Nerve Growth Factor Precursor Is Dependent upon Relative Levels of Its Receptors     
Raheleh Masoudi  Maria S. Ioannou  Michael D. Coughlin  Promila Pagadala  Kenneth E. Neet  Oliver Clewes  Shelley J. Allen  David Dawbarn    Margaret Fahnestock 《The Journal of biological chemistry》2009,284(27):18424-18433
Nerve growth factor (NGF) is produced as a precursor called pro-nerve growth factor (proNGF), which is secreted by many tissues and is the predominant form of NGF in the central nervous system. In Alzheimer disease brain, cholinergic neurons degenerate and can no longer transport NGF as efficiently, leading to an increase in untransported NGF in the target tissue. The protein that accumulates in the target tissue is proNGF, not the mature form. The role of this precursor is controversial, and both neurotrophic and apoptotic activities have been reported for recombinant proNGFs. Differences in the protein structures, protein expression systems, methods used for protein purification, and methods used for bioassay may affect the activity of these proteins. Here, we show that proNGF is neurotrophic regardless of mutations or tags, and no matter how it is purified or in which system it is expressed. However, although proNGF is neurotrophic under our assay conditions for primary sympathetic neurons and for pheochromocytoma (PC12) cells, it is apoptotic for unprimed PC12 cells when they are deprived of serum. The ratio of tropomyosin-related kinase A to p75 neurotrophin receptor is low in unprimed PC12 cells compared with primed PC12 cells and sympathetic neurons, altering the balance of proNGF-induced signaling to favor apoptosis. We conclude that the relative level of proNGF receptors determines whether this precursor exhibits neurotrophic or apoptotic activity.Nerve growth factor (NGF)3 regulates neuronal survival, neurite outgrowth, and differentiation in the peripheral and central nervous systems (1). The mature form of NGF forms a non-covalent homodimer and binds with high affinity (kd ≈ 10−11 m) to tropomyosin-related kinase A (TrkA) and with low affinity (kd ≈ 10−9 m) to the common neurotrophin receptor p75NTR (p75 neurotrophin receptor) (2). NGF promotes cell survival and growth in cells expressing TrkA through activation of the phosphatidylinositol 3-kinase/AKT pathway and the Ras/mitogen-activated protein kinase (MAPK) pathway (3, 4). p75NTR plays diverse roles, ranging from cell survival to cell death depending on the cellular context in which it is expressed. Through activation of the NF-κB pathway, p75NTR can contribute to cell survival in sensory neurons (5), it is involved in axonal growth via regulation of Rho activity (6), and it can interact with Trks to enhance neurotrophin affinity (at low concentration of ligand) and specificity of binding to Trks (79). High levels of p75NTR expression can induce apoptosis when there are low levels of Trk or when Trk is absent (10, 11). Apoptosis occurs through increased ceramide production (12), activation of c-Jun N-terminal kinase (JNK1), and p53 (10, 13). p75NTR requires a co-receptor called sortilin to induce cell death (14).NGF is produced as a precursor called pro-nerve growth factor (proNGF) (15). ProNGF is secreted by many tissues such as prostate cells, spermatids, hair follicles, oral mucosal keratinocytes, sympathetic neurons, cortical astrocytes, heart, and spleen (1620). ProNGF is the predominant form of NGF in the central and peripheral nervous systems, whereas little or no mature NGF can be detected (2124). In Alzheimer disease brain, retrograde transport from the cortex and hippocampus to basal forebrain cholinergic neurons is reduced as these neurons degenerate, with concomitant proNGF accumulation in the cortex and hippocampus (21, 23). This suggested that proNGF mediates biological activity besides its prodomain function of promoting protein folding and regulation of neurotrophin secretion (2528). To study the role of proNGF protein in vitro, point mutations were inserted at the cleavage site used by furin, a proprotein convertase known to cleave proNGF (29), to minimize the conversion of proNGF to mature NGF. The resulting recombinant, cleavage-resistant proNGFs reportedly exhibit either apoptotic activity (30, 31) or neurotrophic activity (32, 33). These recombinant proteins differ in several ways (ProNGF(R−1G)ProNGFhisProNGFEProNGF123WT-NGFhisMutations−1 (R to G)−2 and −1 (RR to AA), 118 and 119 (RR to AA)−1 and +1 (RS to AA)−73 and −72 (RR to AA), −43 and −42 (KKRR to KAAR), −2 and −1 (KR to AA)None: cleavable proNGFTagNo tagHistidine tagNo tagNo tagHistidine tagExpression systemInsect cellsInsect cells, mammalian cellsBacteriaInsect cellsInsect cells, mammalian cellsPurificationNo purificationNickel columnRefolded from inclusion bodies, FPLCCation exchange chromatography, immunoaffinity chromatographyNickel columnOpen in a separate window  相似文献   

12.
Genetic Transformation and Mutagenesis via Single-Stranded DNA in the Unicellular,Diazotrophic Cyanobacteria of the Genus Cyanothece     
Hongtao Min  Louis A. Sherman 《Applied and environmental microbiology》2010,76(22):7641-7645
  相似文献   

13.
Glycopeptide-preferring Polypeptide GalNAc Transferase 10 (ppGalNAc T10), Involved in Mucin-type O-Glycosylation, Has a Unique GalNAc-O-Ser/Thr-binding Site in Its Catalytic Domain Not Found in ppGalNAc T1 or T2     
Cynthia L. Perrine  Anjali Ganguli  Peng Wu  Carolyn R. Bertozzi  Timothy A. Fritz  Jayalakshmi Raman  Lawrence A. Tabak    Thomas A. Gerken 《The Journal of biological chemistry》2009,284(30):20387-20397
Mucin-type O-gly co sy la tion is initiated by a large family of UDP-GalNAc:polypeptide α-N-acetylgalactosaminyltransferases (ppGalNAc Ts) that transfer GalNAc from UDP-GalNAc to the Ser and Thr residues of polypeptide acceptors. Some members of the family prefer previously gly co sylated peptides (ppGalNAc T7 and T10), whereas others are inhibited by neighboring gly co sy la tion (ppGalNAc T1 and T2). Characterizing their peptide and glycopeptide substrate specificity is critical for understanding the biological role and significance of each isoform. Utilizing a series of random peptide and glycopeptide substrates, we have obtained the peptide and glycopeptide specificities of ppGalNAc T10 for comparison with ppGalNAc T1 and T2. For the glycopeptide substrates, ppGalNAc T10 exhibited a single large preference for Ser/Thr-O-GalNAc at the +1 (C-terminal) position relative to the Ser or Thr acceptor site. ppGalNAc T1 and T2 revealed no significant enhancements suggesting Ser/Thr-O-GalNAc was inhibitory at most positions for these isoforms. Against random peptide substrates, ppGalNAc T10 revealed no significant hydrophobic or hydrophilic residue enhancements, in contrast to what has been reported previously for ppGalNAc T1 and T2. Our results reveal that these transferases have unique peptide and glycopeptide preferences demonstrating their substrate diversity and their likely roles ranging from initiating transferases to filling-in transferases.Mucin-type O-glycosylation is a common post-translational modification of secreted and membrane-associated proteins. O-Glycan biosynthesis is initiated by the transfer of GalNAc from UDP-GalNAc to the hydroxyl groups of serine or threonine residues in a polypeptide, catalyzed by a family of polypeptide N-α-acetylgalactosaminyltransferases (ppGalNAc Ts).5 To date, 16 mammalian members have been reported in the literature (116) with a total of at least 20 members currently present in the human genome data base. Multiple members of the ppGalNAc T family have also been identified in Drosophila (9, 10, 14), Caenorhabditis elegans (3, 8), and single and multicellular organisms (1720). Several members show close sequence orthologues across species suggesting that the ppGalNAc Ts are responsible for biologically significant functions that have been conserved during evolution. For example, in Drosophila four isoforms have close sequence orthologues to the mammalian transferases. Of the two that have been recently compared, nearly identical peptide substrate specificities have been observed between the fly and mammals, suggesting common but presently unknown functions preserved across these diverse species (21).Recently, several ppGalNAc T isoforms have been shown to be important for normal development or cellular processes. For example, inactive mutations in the fly PGANT35A (the T11 orthologue in mammals) are lethal because of the disruption of the tracheal tube structures (9, 10, 22), whereas mutations in PGANT3 alter epithelial cell adhesion in the Drosophila wing blade resulting in wing blistering (23). In humans, mutations in ppGalNAc T3 are associated with familial tumoral calcinosis, the result of the abnormal processing and secretion of the phosphaturic factor FGF23 (24, 25). Human ppGalNAc T14 has been suggested to modulate apoptotic signaling in tumor cells by its glycosylation of the proapoptotic receptors DLR4 and DLR5 (26), and very recently the specific O-glycosylation of the TGFB-II receptor (ActR-II) by the GalNTL1 has been shown to modulate its signaling in development (16).Historically, the major targets of the ppGalNAc Ts have been thought to be heavily O-glycosylated mucin domains of membrane and secreted glycoproteins. Such domains typically contain 15–30% Ser or Thr, which are highly (>50%) substituted by GalNAc. One question in the field is as follows. How is this high degree of peptide core glycosylation achieved and is it related to the large number of ppGalNAc isoforms, some of which may even have specific mucin domain preferences? Interestingly, some members of the ppGalNAc T family are known to prefer substrates that have been previously modified with O-linked GalNAc on nearby Ser/Thr residues, hence having so-called glycopeptide or filling-in activities, i.e. ppGalNAc T7 and T10 (8, 2729). Others simply possess altered preferences against glycopeptide substrates, i.e. ppGalNAc T2 and T4 (3033), or may be inhibited by neighboring glycosylation, i.e. ppGalNAc T1 and T2 (29, 34, 35). These latter transferases have been called early or initiating transferases, preferring nonglycosylated over-glycosylated substrates. Presently, little is known about which factors dictate the different peptide/glycopeptide specificities among the ppGalNAc Ts.The ppGalNAc Ts consist of an N-terminal catalytic domain tethered by a short linker to a C-terminal ricin-like lectin domain containing three recognizable carbohydrate-binding sites (36). Because ppGalNAc T7 and T10 prefer to transfer GalNAc to glycopeptide acceptors, it has been widely assumed that their C-terminal lectin domains would play significant roles in this activity, as has been demonstrated for other family members (27, 28, 32). Recently, Kubota et al. (37) solved the crystal structure of ppGalNAc T10 in complex with Ser-GalNAc specifically bound to its lectin domain. In this work (37), the authors further demonstrated that a T10 lectin domain mutant indeed had altered specificity against GalNAc-containing glycopeptide substrates when the acceptor Ser/Thr site was distal from the pre-existing glycopeptide GalNAc site. However, it was also observed that the lectin mutant still possessed relatively unaltered glycopeptide activity when the acceptor Ser/Thr site was directly N-terminal of a pre-existing glycopeptide GalNAc site. Kubota et al. (37) therefore concluded that for ppGalNAc T10, both its lectin and indeed its catalytic domain must contain distinct peptide GalNAc recognition sites. In support of this, Raman et al. (33) have shown that the complete removal of the ppGalNAc T10 lectin domain only slightly alters its specificity against distal glycopeptide substrates while showing no difference in its ability to glycosylate residues directly N-terminal of an existing site of glycosylation. Thus, it seems that the catalytic domain of ppGalNAc T10 may have specific requirements for a peptide O-linked GalNAc in at least the +1 position (toward the C terminus) of residues being glycosylated. As no systematic determination of the glycopeptide binding properties of the ppGalNAc Ts catalytic domain has been performed, it is unknown whether additional GalNAc peptide-binding sites exist in T10 or, for that matter, any of the other ppGalNAc Ts.We have recently reported the use of oriented random peptide substrates, GAGA(X)nT(X)nAGAGK (where X indicates randomized amino acid positions and n = 3 and 5) for determining the peptide substrate specificities of mammalian ppGalNAc T1, T2, and their fly orthologues (21, 38). In the present work, we extend this approach to the determination of the catalytic domain glycopeptide (Ser/Thr-O-GalNAc) substrate preferences for ppGalNAc T1, T2, and T10 employing two n = 4 oriented random glycopeptide libraries (21). Interestingly, ppGalNAc T10 displays few significant enhancements and specifically lacks the Pro residue enhancements observed for ppGalNAc T1 and T2. These findings further demonstrate the vast substrate diversity of the catalytic domains of the ppGalNAc T family of transferases.

TABLE 1

ppGalNAc transferase random substrates utilized in this workPVI, PVII, GP-I, and GP-II random (glyco)peptide substrates.
PeptideSequenceNo. of unique sequences
GAGAXXXXXTXXXXXAGAGK
P-VIX = G, A, P, V, L, Y, E, Q, R, H10 × 109
P-VIIX = G, A, P, I, M, F, D, N, R, K10 × 109

GAGAXXXXTXXXXAGAG
GP-IX = G, A, P, V, I, F, Y, E, D, N, R, K, H, and Ser-O-α-GalNAc1.47 × 109

GAGAXXXX(Thr-O-α-GalNAc)XXXXAGAG
GP-IIX = G, A, P, V, I, F, Y, E, D, N, R, K, H, S1.47 × 109
Open in a separate window  相似文献   

14.
Presence and Prevalence of Viruses in Local and Migratory Honeybees (Apis mellifera) in Massachusetts   总被引:1,自引:0,他引:1  
Anna Welch  Francis Drummond  Sunil Tewari  Anne Averill  John P. Burand 《Applied and environmental microbiology》2009,75(24):7862-7865
  相似文献   

15.
Normalization and Statistical Analysis of Multiplexed Bead-based Immunoassay Data Using Mixed-effects Modeling     
David C. Clarke  Melody K. Morris  Douglas A. Lauffenburger 《Molecular & cellular proteomics : MCP》2013,12(1):245-262
  相似文献   

16.
Focus on Chromatin/Epigenetics: Trans-Homolog Interactions Facilitating Paramutation in Maize     
Brian John Giacopelli  Jay Brian Hollick 《Plant physiology》2015,168(4):1226-1236
  相似文献   

17.
Neurodegeneration and Alzheimer's disease (AD). What Can Proteomics Tell Us About the Alzheimer's Brain?     
Guillermo Moya-Alvarado  Noga Gershoni-Emek  Eran Perlson  Francisca C. Bronfman 《Molecular & cellular proteomics : MCP》2016,15(2):409-425
  相似文献   

18.
Natural Infection of Burkholderia pseudomallei in an Imported Pigtail Macaque (Macaca nemestrina) and Management of the Exposed Colony     
Crystal H Johnson  Brianna L Skinner  Sharon M Dietz  David Blaney  Robyn M Engel  George W Lathrop  Alex R Hoffmaster  Jay E Gee  Mindy G Elrod  Nathaniel Powell  Henry Walke 《Comparative medicine》2013,63(6):528-535
Identification of the select agent Burkholderia pseudomallei in macaques imported into the United States is rare. A purpose-bred, 4.5-y-old pigtail macaque (Macaca nemestrina) imported from Southeast Asia was received from a commercial vendor at our facility in March 2012. After the initial acclimation period of 5 to 7 d, physical examination of the macaque revealed a subcutaneous abscess that surrounded the right stifle joint. The wound was treated and resolved over 3 mo. In August 2012, 2 mo after the stifle joint wound resolved, the macaque exhibited neurologic clinical signs. Postmortem microbiologic analysis revealed that the macaque was infected with B. pseudomallei. This case report describes the clinical evaluation of a B. pseudomallei-infected macaque, management and care of the potentially exposed colony of animals, and protocols established for the animal care staff that worked with the infected macaque and potentially exposed colony. This article also provides relevant information on addressing matters related to regulatory issues and risk management of potentially exposed animals and animal care staff.Abbreviations: CDC, Centers for Disease Control and Prevention; IHA, indirect hemagglutination assay; PEP, postexposure prophylacticBurkholderia pseudomallei, formerly known as Pseudomonas pseudomallei, is a gram-negative, aerobic, bipolar, motile, rod-shaped bacterium. B. pseudomallei infections (melioidosis) can be severe and even fatal in both humans and animals. This environmental saprophyte is endemic to Southeast Asia and northern Australia, but it has also been found in other tropical and subtropical areas of the world.7,22,32,42 The bacterium is usually found in soil and water in endemic areas and is transmitted to humans and animals primarily through percutaneous inoculation, ingestion, or inhalation of a contaminated source.8, 22,28,32,42 Human-to-human, animal-to-animal, and animal-to-human spread are rare.8,32 In December 2012, the National Select Agent Registry designated B. pseudomallei as a Tier 1 overlap select agent.39 Organisms classified as Tier 1 agents present the highest risk of deliberate misuse, with the most significant potential for mass casualties or devastating effects to the economy, critical infrastructure, or public confidence. Select agents with this status have the potential to pose a severe threat to human and animal health or safety or the ability to be used as a biologic weapon.39Melioidosis in humans can be challenging to diagnose and treat because the organism can remain latent for years and is resistant to many antibiotics.12,37,41 B. pseudomallei can survive in phagocytic cells, a phenomenon that may be associated with latent infections.19,38 The incubation period in naturally infected animals ranges from 1 d to many years, but symptoms typically appear 2 to 4 wk after exposure.13,17,35,38 Disease generally presents in 1 of 2 forms: localized infection or septicemia.22 Multiple methods are used to diagnose melioidosis, including immunofluorescence, serology, and PCR analysis, but isolation of the bacteria from blood, urine, sputum, throat swabs, abscesses, skin, or tissue lesions remains the ‘gold standard.’9,22,40,42 The prognosis varies based on presentation, time to diagnosis, initiation of appropriate antimicrobial treatment, and underlying comorbidities.7,28,42 Currently, there is no licensed vaccine to prevent melioidosis.There are several published reports of naturally occurring melioidosis in a variety of nonhuman primates (NHP; 2,10,13,17,25,30,31,35 The first reported case of melioidosis in monkeys was recorded in 1932, and the first published case in a macaque species was in 1966.30 In the United States, there have only been 7 documented cases of NHP with B. pseudomallei infection.2,13,17 All of these cases occurred prior to the classification of B. pseudomallei as a select agent. Clinical signs in NHP range from subclinical or subacute illness to acute septicemia, localized infection, and chronic infection. NHP with melioidosis can be asymptomatic or exhibit clinical signs such as anorexia, wasting, purulent drainage, subcutaneous abscesses, and other soft tissue lesions. Lymphadenitis, lameness, osteomyelitis, paralysis and other CNS signs have also been reported.2,7,10,22,28,32 In comparison, human''s clinical signs range from abscesses, skin ulceration, fever, headache, joint pain, and muscle tenderness to abdominal pain, anorexia, respiratory distress, seizures, and septicemia.7,9,21,22

Table 1.

Summary of reported cases of naturally occurring Burkholderia pseudomalleiinfections in nonhuman primates
CountryaImported fromDate reportedSpeciesReference
AustraliaBorneo1963Pongo sp.36
BruneiUnknown1982Orangutan (Pongo pygmaeus)33
France1976Hamlyn monkey (Cercopithecus hamlyni) Patas monkey (Erythrocebus patas)11
Great BritainPhilippines and Indonesia1992Cynomolgus monkey (Macaca fascicularis)10
38
MalaysiaUnknown1966Macaca spp.30
Unknown1968Spider monkey (Brachytelis arachnoides) Lar gibbon (Hylobates lar)20
Unknown1969Pig-tailed macaque (Macaca nemestrina)35
Unknown1984Banded leaf monkey (Presbytis melalophos)25
SingaporeUnknown1995Gorillas, gibbon, mandrill, chimpanzee43
ThailandUnknown2012Monkey19
United StatesThailand1970Stump-tailed macaque (Macaca arctoides)17
IndiaPig-tailed macaque (Macaca nemestrina)
AfricaRhesus macaque (Macaca mulatta) Chimpanzee (Pan troglodytes)
Unknown1971Chimpanzee (Pan troglodytes)3
Malaysia1981Pig-tailed macaque (Macaca nemestrina)2
Wild-caught, unknown1986Rhesus macaque (Macaca mulatta)13
Indonesia2013Pig-tailed macaque (Macaca nemestrina)Current article
Open in a separate windowaCountry reflects the location where the animal was housed at the time of diagosis.Here we describe a case of melioidosis diagnosed in a pigtail macaque (Macaca nemestrina) imported into the United States from Indonesia and the implications of the detection of a select agent identified in a laboratory research colony. We also discuss the management and care of the exposed colony, zoonotic concerns regarding the animal care staff that worked with the shipment of macaques, effects on research studies, and the procedures involved in reporting a select agent incident.  相似文献   

19.
RNA Polymerase I Transcription Silences Noncoding RNAs at the Ribosomal DNA Locus in Saccharomyces cerevisiae     
Elisa Cesarini  Francesca Romana Mariotti  Francesco Cioci  Giorgio Camilloni 《Eukaryotic cell》2010,9(2):325-335
  相似文献   

20.
Proteomics of Saccharomyces cerevisiae Organelles     
Elena Wiederhold  Liesbeth M. Veenhoff  Bert Poolman    Dirk Jan Slotboom 《Molecular & cellular proteomics : MCP》2010,9(3):431-445
  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号