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.
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   

2.
A Boolean network is a model used to study the interactions between different genes in genetic regulatory networks. In this paper, we present several algorithms using gene ordering and feedback vertex sets to identify singleton attractors and small attractors in Boolean networks. We analyze the average case time complexities of some of the proposed algorithms. For instance, it is shown that the outdegree-based ordering algorithm for finding singleton attractors works in time for , which is much faster than the naive time algorithm, where is the number of genes and is the maximum indegree. We performed extensive computational experiments on these algorithms, which resulted in good agreement with theoretical results. In contrast, we give a simple and complete proof for showing that finding an attractor with the shortest period is NP-hard.[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   

3.
4.
5.
Ribosomal protein S1 has been shown to be a significant effector of prokaryotic translation. The protein is in fact capable of efficiently initiating translation, regardless of the presence of a Shine-Dalgarno sequence in mRNA. Structural insights into this process have remained elusive, as S1 is recalcitrant to traditional techniques of structural analysis, such as x-ray crystallography. Through the application of protein cross-linking and high resolution mass spectrometry, we have detailed the ribosomal binding site of S1 and have observed evidence of its dynamics. Our results support a previous hypothesis that S1 acts as the mRNA catching arm of the prokaryotic ribosome. We also demonstrate that in solution the major domains of the 30S subunit are remarkably flexible, capable of moving 30–50Å with respect to one another.Initiation of translation is often the rate-limiting step of protein biosynthesis (1). In prokaryotes, this process is widely recognized to be directed by the Shine-Dalgarno (S.D.)1 sequence of mRNA and its complementation with the 3′ end of 16S rRNA (2). However, binding of the S.D. sequence to the ribosome is not obligatory for initiation. Ribosomal protein S1, widely conserved in prokaryotes, (3) has been shown to efficiently initiate translation, regardless of the presence of an S.D. sequence (4, 5).S1 is a strikingly atyptical ribosomal protein, being both the largest (61 kDa) and the most acidic (pI 4.7) (6). The protein is composed of six homologous repeats each forming beta barrel domains (3) that in solution comprise a highly elongated structure spanning up to ca. 230 Å (7). This length is comparable to the diameter of the ribosome itself. In addition to these anomalous characteristics, S1 is also one of only two ribosomal proteins that has been attributed functional significance (6). Ribosomal protein S1, for instance, has no apparent role in the assembly of the ribosome, (2) yet is critical for translation in E. coli (8, 9). The functional significance of S1 is related to its most pronounced characteristic, the ability to simultaneously bind mRNA and the ribosome. Analysis of fragments produced by limited proteolysis and chemical cleavage of S1 has shown that an N-terminal fragment of S1 (residues 1–193) binds the ribosome (10) but not RNA (11). Likewise, a C-terminal fragment (res 172–557) binds RNA (12, 13) but not the ribosome (6, 10). By nature of this bi-functional structure, S1 enhances the E. coli ribosome''s affinity for RNA ∼5000 fold (14) and can directly mediate initiation of translation by binding the 5′ UTR of mRNA (4, 5). These observations have led to the hypothesis that S1 acts as a catching arm for the prokaryotic ribosome, working to bring mRNA to the proximity of the ribosome and thereby facilitate initiation (6).Unfortunately, structural analyses capturing how S1 is able to function in this manner remain elusive. A high-resolution crystal structure of ribosome bound S1, or even free S1, does not exist, because S1 is recalcitrant to crystallography (6). Preparation of ribosomes for x-ray crystallography actually involves the deliberate removal of ribosomal protein S1 as a means to improve the reproducibility of crystallization and the quality of the ribosome crystals formed (1517). The structure and interactions of the protein have nevertheless intrigued structural biologists for decades. However, studies completed to date have failed to convincingly demonstrate the interaction between S1 and the rest of the 30S subunit, because they were incapable of localizing the individual S1 domains (16, 1820).We have studied the binding of S1 to the 30S subunit by combining cross-linking with mass spectrometry. Chemical cross-linking has long been appreciated as a technique to probe protein-protein interactions (21, 22). With the advent of modern mass spectrometers, it can be very effectively employed to confidently identify the exact residues involved in linkages (2328). In most cross-linking analyses, protein residues are targeted for covalent modification with a molecule that contains two reactive groups separated by a spacer arm of known length. Only protein residues closer than the length of the spacer arm are capable of being linked. Identification of cross-linked residues thereby provides distance constraints for structural modeling. In this work, the novel amidinating protein cross-linker, DEST (diethyl suberthioimidate), was employed (29, 30). This amine reactive reagent, unlike commercially available reagents, preserves the native basicity of the residues it modifies while being effective at physiological pH. Use of the reagent is unlikely to perturb protein structure and the modifications it imparts are compatible with ionization for mass spectrometry. We have additionally shown that the cross-links it forms can be efficiently enriched from other components of proteolytic digests using strong cation exchange (SCX) chromatography, (30) and that DEST cross-linking of ribosomes yields structural information in excellent agreement with x-ray crystallography (29). Although DEST is an 11Å spacer arm cross-linker, it links alpha carbons up to 24Å apart because of the length and flexibility of lysine side chains. Nevertheless, this is sufficient resolution to approximate the binding positions of the 10kDa domains of S1. Furthermore, multiple cross-linking of a single domain significantly enhances the resolution with which it can be localized.Here, through the application of protein cross-linking and high resolution mass spectrometry, we show that S1 binds to the 30S subunit near the anti-S.D. motif of the 16S rRNA, demonstrate that it is highly elongated even when bound to the ribosome, and provide evidence that its C-terminal mRNA binding region is remarkably dynamic. Our results thus indicate S1 is structurally poised, as previously hypothesized, (6) to act as the mRNA catching arm of the prokaryotic ribosome.  相似文献   

6.
7.
Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.Insulin plays a central role in the regulation of vertebrate metabolism. The hormone, the post-translational product of a single-chain precursor, is a globular protein containing two chains, A (21 residues) and B (30 residues). Recent advances in human genetics have identified dominant mutations in the insulin gene causing permanent neonatal-onset DM2 (14). The mutations are predicted to block folding of the precursor in the ER of pancreatic β-cells. Although expression of the wild-type allele would in other circumstances be sufficient to maintain homeostasis, studies of a corresponding mouse model (57) suggest that the misfolded variant perturbs wild-type biosynthesis (8, 9). Impaired β-cell secretion is associated with ER stress, distorted organelle architecture, and cell death (10). These findings have renewed interest in insulin biosynthesis (1113) and the structural basis of disulfide pairing (1419). Protein evolution is constrained not only by structure and function but also by susceptibility to toxic misfolding.  相似文献   

8.
A variety of high-throughput methods have made it possible to generate detailed temporal expression data for a single gene or large numbers of genes. Common methods for analysis of these large data sets can be problematic. One challenge is the comparison of temporal expression data obtained from different growth conditions where the patterns of expression may be shifted in time. We propose the use of wavelet analysis to transform the data obtained under different growth conditions to permit comparison of expression patterns from experiments that have time shifts or delays. We demonstrate this approach using detailed temporal data for a single bacterial gene obtained under 72 different growth conditions. This general strategy can be applied in the analysis of data sets of thousands of genes under different conditions.[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   

9.
10.
11.
12.
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   

13.
14.
Decomposing a biological sequence into its functional regions is an important prerequisite to understand the molecule. Using the multiple alignments of the sequences, we evaluate a segmentation based on the type of statistical variation pattern from each of the aligned sites. To describe such a more general pattern, we introduce multipattern consensus regions as segmented regions based on conserved as well as interdependent patterns. Thus the proposed consensus region considers patterns that are statistically significant and extends a local neighborhood. To show its relevance in protein sequence analysis, a cancer suppressor gene called p53 is examined. The results show significant associations between the detected regions and tendency of mutations, location on the 3D structure, and cancer hereditable factors that can be inferred from human twin studies.[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]  相似文献   

15.
The Dbf4-Cdc7 kinase (DDK) is required for the activation of the origins of replication, and DDK phosphorylates Mcm2 in vitro. We find that budding yeast Cdc7 alone exists in solution as a weakly active multimer. Dbf4 forms a likely heterodimer with Cdc7, and this species phosphorylates Mcm2 with substantially higher specific activity. Dbf4 alone binds tightly to Mcm2, whereas Cdc7 alone binds weakly to Mcm2, suggesting that Dbf4 recruits Cdc7 to phosphorylate Mcm2. DDK phosphorylates two serine residues of Mcm2 near the N terminus of the protein, Ser-164 and Ser-170. Expression of mcm2-S170A is lethal to yeast cells that lack endogenous MCM2 (mcm2Δ); however, this lethality is rescued in cells harboring the DDK bypass mutant mcm5-bob1. We conclude that DDK phosphorylation of Mcm2 is required for cell growth.The Cdc7 protein kinase is required throughout the yeast S phase to activate origins (1, 2). The S phase cyclin-dependent kinase also activates yeast origins of replication (35). It has been proposed that Dbf4 activates Cdc7 kinase in S phase, and that Dbf4 interaction with Cdc7 is essential for Cdc7 kinase activity (6). However, it is not known how Dbf4-Cdc7 (DDK)2 acts during S phase to trigger the initiation of DNA replication. DDK has homologs in other eukaryotic species, and the role of Cdc7 in activation of replication origins during S phase may be conserved (710).The Mcm2-7 complex functions with Cdc45 and GINS to unwind DNA at a replication fork (1115). A mutation of MCM5 (mcm5-bob1) bypasses the cellular requirements for DBF4 and CDC7 (16), suggesting a critical physiologic interaction between Dbf4-Cdc7 and Mcm proteins. DDK phosphorylates Mcm2 in vitro with proteins purified from budding yeast (17, 18) or human cells (19). Furthermore, there are mutants of MCM2 that show synthetic lethality with DBF4 mutants (6, 17), suggesting a biologically relevant interaction between DBF4 and MCM2. Nevertheless, the physiologic role of DDK phosphorylation of Mcm2 is a matter of dispute. In human cells, replacement of MCM2 DDK-phosphoacceptor residues with alanines inhibits DNA replication, suggesting that Dbf4-Cdc7 phosphorylation of Mcm2 in humans is important for DNA replication (20). In contrast, mutation of putative DDK phosphorylation sites at the N terminus of Schizosaccharomyces pombe Mcm2 results in viable cells, suggesting that phosphorylation of S. pombe Mcm2 by DDK is not critical for cell growth (10).In budding yeast, Cdc7 is present at high levels in G1 and S phase, whereas Dbf4 levels peak in S phase (18, 21, 22). Furthermore, budding yeast DDK binds to chromatin during S phase (6), and it has been shown that Dbf4 is required for Cdc7 binding to chromatin in budding yeast (23, 24), fission yeast (25), and Xenopus (9). Human and fission yeast Cdc7 are inert on their own (7, 8), but Dbf4-Cdc7 is active in phosphorylating Mcm proteins in budding yeast (6, 26), fission yeast (7), and human (8, 10). Based on these data, it has been proposed that Dbf4 activates Cdc7 kinase in S phase and that Dbf4 interaction with Cdc7 is essential for Cdc7 kinase activity (6, 9, 18, 2124). However, a mechanistic analysis of how Dbf4 activates Cdc7 has not yet been accomplished. For example, the multimeric state of the active Dbf4-Cdc7 complex is currently disputed. A heterodimer of fission yeast Cdc7 (Hsk1) in complex with fission yeast Dbf4 (Dfp1) can phosphorylate Mcm2 (7). However, in budding yeast, oligomers of Cdc7 exist in the cell (27), and Dbf4-Cdc7 exists as oligomers of 180 and 300 kDa (27).DDK phosphorylates the N termini of human Mcm2 (19, 20, 28), human Mcm4 (10), budding yeast Mcm4 (26), and fission yeast Mcm6 (10). Although the sequences of the Mcm N termini are poorly conserved, the DDK sites identified in each study have neighboring acidic residues. The residues of budding yeast Mcm2 that are phosphorylated by DDK have not yet been identified.In this study, we find that budding yeast Cdc7 is weakly active as a multimer in phosphorylating Mcm2. However, a low molecular weight form of Dbf4-Cdc7, likely a heterodimer, has a higher specific activity for phosphorylation of Mcm2. Dbf4 or DDK, but not Cdc7, binds tightly to Mcm2, suggesting that Dbf4 recruits Cdc7 to Mcm2. DDK phosphorylates two serine residues of Mcm2, Ser-164 and Ser-170, in an acidic region of the protein. Mutation of Ser-170 is lethal to yeast cells, but this phenotype is rescued by the DDK bypass mutant mcm5-bob1. We conclude that DDK phosphorylation of Ser-170 of Mcm2 is required for budding yeast growth.  相似文献   

16.
Previous studies have shown that protein-protein interactions among splicing factors may play an important role in pre-mRNA splicing. We report here identification and functional characterization of a new splicing factor, Sip1 (SC35-interacting protein 1). Sip1 was initially identified by virtue of its interaction with SC35, a splicing factor of the SR family. Sip1 interacts with not only several SR proteins but also with U1-70K and U2AF65, proteins associated with 5′ and 3′ splice sites, respectively. The predicted Sip1 sequence contains an arginine-serine-rich (RS) domain but does not have any known RNA-binding motifs, indicating that it is not a member of the SR family. Sip1 also contains a region with weak sequence similarity to the Drosophila splicing regulator suppressor of white apricot (SWAP). An essential role for Sip1 in pre-mRNA splicing was suggested by the observation that anti-Sip1 antibodies depleted splicing activity from HeLa nuclear extract. Purified recombinant Sip1 protein, but not other RS domain-containing proteins such as SC35, ASF/SF2, and U2AF65, restored the splicing activity of the Sip1-immunodepleted extract. Addition of U2AF65 protein further enhanced the splicing reconstitution by the Sip1 protein. Deficiency in the formation of both A and B splicing complexes in the Sip1-depleted nuclear extract indicates an important role of Sip1 in spliceosome assembly. Together, these results demonstrate that Sip1 is a novel RS domain-containing protein required for pre-mRNA splicing and that the functional role of Sip1 in splicing is distinct from those of known RS domain-containing splicing factors.Pre-mRNA splicing takes place in spliceosomes, the large RNA-protein complexes containing pre-mRNA, U1, U2, U4/6, and U5 small nuclear ribonucleoprotein particles (snRNPs), and a large number of accessory protein factors (for reviews, see references 21, 22, 37, 44, and 48). It is increasingly clear that the protein factors are important for pre-mRNA splicing and that studies of these factors are essential for further understanding of molecular mechanisms of pre-mRNA splicing.Most mammalian splicing factors have been identified by biochemical fractionation and purification (3, 15, 19, 3136, 45, 6971, 73), by using antibodies recognizing splicing factors (8, 9, 16, 17, 61, 66, 67, 74), and by sequence homology (25, 52, 74).Splicing factors containing arginine-serine-rich (RS) domains have emerged as important players in pre-mRNA splicing. These include members of the SR family, both subunits of U2 auxiliary factor (U2AF), and the U1 snRNP protein U1-70K (for reviews, see references 18, 41, and 59). Drosophila alternative splicing regulators transformer (Tra), transformer 2 (Tra2), and suppressor of white apricot (SWAP) also contain RS domains (20, 40, 42). RS domains in these proteins play important roles in pre-mRNA splicing (7, 71, 75), in nuclear localization of these splicing proteins (23, 40), and in protein-RNA interactions (56, 60, 64). Previous studies by us and others have demonstrated that one mechanism whereby SR proteins function in splicing is to mediate specific protein-protein interactions among spliceosomal components and between general splicing factors and alternative splicing regulators (1, 1a, 6, 10, 27, 63, 74, 77). Such protein-protein interactions may play critical roles in splice site recognition and association (for reviews, see references 4, 18, 37, 41, 47 and 59). Specific interactions among the splicing factors also suggest that it is possible to identify new splicing factors by their interactions with known splicing factors.Here we report identification of a new splicing factor, Sip1, by its interaction with the essential splicing factor SC35. The predicted Sip1 protein sequence contains an RS domain and a region with sequence similarity to the Drosophila splicing regulator, SWAP. We have expressed and purified recombinant Sip1 protein and raised polyclonal antibodies against the recombinant Sip1 protein. The anti-Sip1 antibodies specifically recognize a protein migrating at a molecular mass of approximately 210 kDa in HeLa nuclear extract. The anti-Sip1 antibodies sufficiently deplete Sip1 protein from the nuclear extract, and the Sip1-depleted extract is inactive in pre-mRNA splicing. Addition of recombinant Sip1 protein can partially restore splicing activity to the Sip1-depleted nuclear extract, indicating an essential role of Sip1 in pre-mRNA splicing. Other RS domain-containing proteins, including SC35, ASF/SF2, and U2AF65, cannot substitute for Sip1 in reconstituting splicing activity of the Sip1-depleted nuclear extract. However, addition of U2AF65 further increases splicing activity of Sip1-reconstituted nuclear extract, suggesting that there may be a functional interaction between Sip1 and U2AF65 in nuclear extract.  相似文献   

17.
18.
19.
20.
In Archaea, an hexameric ATPase complex termed PAN promotes proteins unfolding and translocation into the 20 S proteasome. PAN is highly homologous to the six ATPases of the eukaryotic 19 S proteasome regulatory complex. Thus, insight into the mechanism of PAN function may reveal a general mode of action mutual to the eukaryotic 19 S proteasome regulatory complex. In this study we generated a three-dimensional model of PAN from tomographic reconstruction of negatively stained particles. Surprisingly, this reconstruction indicated that the hexameric complex assumes a two-ring structure enclosing a large cavity. Assessment of distinct three-dimensional functional states of PAN in the presence of adenosine 5′-O-(thiotriphosphate) and ADP and in the absence of nucleotides outlined a possible mechanism linking nucleotide binding and hydrolysis to substrate recognition, unfolding, and translocation. A novel feature of the ATPase complex revealed in this study is a gate controlling the “exit port” of the regulatory complex and, presumably, translocation into the 20 S proteasome. Based on our structural and biochemical findings, we propose a possible model in which substrate binding and unfolding are linked to structural transitions driven by nucleotide binding and hydrolysis, whereas translocation into the proteasome only depends upon the presence of an unfolded substrate and binding but not hydrolysis of nucleotide.In eukaryotic cells most protein breakdown in the cytosol and nucleus is catalyzed by the 26 S proteasome. This ∼2.5-MDa (1) complex degrades ubiquitin-conjugated and certain non-ubiquitinated proteins in an ATP-dependent manner (2, 3). The 26 S complex is composed of one or two 19 S regulatory particles situated at the ends of the cylindrical 20 S proteasome. Within the 26 S complex, proteins are hydrolyzed in the 20 S proteasome. Tagged substrates, however, first bind to the 19 S regulatory particle, which catalyzes their unfolding and translocation into the 20 S subcomplex (4, 5). The 19 S regulatory particle consists of at least 17 different subunits (1, 6). Nine of these subunits form a “lid,” whereas the other eight subunits, including six ATPases, comprise the base of the 19 S particle. Electron microscopy (710) as well as cross-linking experiments (11, 12) have demonstrated that the six homologous ATPases are associated with the α rings of the 20 S particle.Unlike eukaryotes, Archaea and certain eubacteria contain homologous 20 S particles but lack ubiquitin. Their proteasomes degrade proteins in association with a hexameric ATPase ring complex termed PAN (13). PAN appears to be the evolutionary precursor of the 19 S base, predating the coupling of ubiquitination and proteolysis in eukaryotes (14). In addition, PAN recognizes the bacterial targeting sequence ssrA (in analogy to the polyubiquitin conjugates in eukaryotes) and efficiently unfolds and translocates globular substrates, like green fluorescent protein, when tagged with ssrA (15). In both PAN and the 19 S proteasome regulatory complexes, ATP is essential for substrate unfolding and translocation and for opening of the gated channel in the α ring through which substrates enter the 20 S particle (1517). Because this portal is quite narrow (1820), only extended polypeptides can enter the 20 S proteasome. Consequently, a globular substrate must be unfolded by the associated ATPase complex to be translocated and digested within the 20 S particle.PAN and the six ATPases found at the base of the 19 S particle are members of the AAA+ superfamily of multimeric ATPases which also includes the ATP-dependent proteases Lon and FtsH and the regulatory components of the bacterial ATP-dependent proteases ClpAP, ClpXP, and HslUV (8, 21). For mechanistic studies of the roles of ATP, the simpler archaeal PAN-20 S system offers many technical advantages over the much more complex 26 S proteasome. For example, prior studies of PAN (17, 22) demonstrated that unfolding of globular substrates (e.g. green fluorescent protein-ssrA) requires ATP hydrolysis. The same was also shown for the Escherichia coli ATP-dependent proteases ClpXP (23) and ClpAP (24). We have also shown that unfolding by PAN can take place on the surface of the ATPase ring in the absence of translocation (15). Thus, unfolding seems to proceed independently from protein translocation into the 20 S proteolytic particle. It is noteworthy that other studies suggest that proteins are unfolded by energy-dependent translocation through the ATPase ring (25, 26). These studies have suggested that the translocation of an unfolded polypeptide from the ATPase into the 20 S core is an active process that is coupled to ATP hydrolysis. A key to underline a detailed molecular mechanism for substrate binding, unfolding, and translocation by the proteasome regulatory ATPase complex is improved understanding of its architecture and the nucleotide-dependent structural transitions that afford these functions.To date we and others have failed to generate micrographs suitable for three-dimensional reconstruction of PAN using single-particle EM analysis. Likewise, structural information regarding the three-dimensional architecture and subunit organization within the 19 S particle is very limited. In fact, high resolution three-dimensional information on the 19 S complex is not yet available. Most knowledge available is based on cross-linking experiments (11, 12) as well as EM structural analysis (710), which provided a three-dimensional model outline of the general architecture of the 26 S complex. Unlike the 19 S complex, the structure of the 20 S subcomplex was determined by x-ray crystallography (18, 19). In contrast to the highly homogenous structure of the 20 S complex, the structural heterogeneity and flexibility of the 19 S subcomplex is presumably reflected in multiple conformations, which in turn also contribute to the difficulty in generating a high resolution three-dimensional structural model of the 26 S proteasome. Accordingly, the initial goal of this study was to generate a three-dimensional model of PAN that will allow us to determine its general architecture and to correlate unique conformational transitions within this ATPase with the nucleotide state of the complex (i.e. in the presence of ATPγS, ADP, or in the absence of nucleotides).Smith et al. (27) suggested a general architecture for the PAN-20 S complex based on two-dimensional averaging of a Thermoplasma acidophilum (TA)3 20 S proteasome and Methanococcus jannaschii (MJ) PAN hybrid complex in the presence of ATPγS. Based on side-view projections of that complex, these authors proposed that PAN assumes an overall structure similar to E. coli HslU (2830).We realized that although PAN appears heterogeneous in electron micrographs, it does not occupy all possible orientations when adsorbed to carbon-coated electron microscopy (EM) grids, a prerequisite for single particle analysis. This problem was overcome by applying electron tomography in conjunction with a three-dimensional averaging procedure that accounts for the missing wedge in the Fourier space of electron tomograms (31, 32). The three-dimensional model generated revealed an unexpected architecture leading to a possible molecular mechanism describing the function of PAN and presumably the 19 S ATPases.  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