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.
Upon completion of synthesis of an Okazaki fragment, the lagging strand replicase must recycle to the next primer at the replication fork in under 0.1 s to sustain the physiological rate of DNA synthesis. We tested the collision model that posits that cycling is triggered by the polymerase encountering the 5′-end of the preceding Okazaki fragment. Probing with surface plasmon resonance, DNA polymerase III holoenzyme initiation complexes were formed on an immobilized gapped template. Initiation complexes exhibit a half-life of dissociation of approximately 15 min. Reduction in gap size to 1 nt increased the rate of dissociation 2.5-fold, and complete filling of the gap increased the off-rate an additional 3-fold (t1/2  ∼ 2 min). An exogenous primed template and ATP accelerated dissociation an additional 4-fold in a reaction that required complete filling of the gap. Neither a 5′-triphosphate nor a 5′-RNA terminated oligonucleotide downstream of the polymerase accelerated dissociation further. Thus, the rate of polymerase release upon gap completion and collision with a downstream Okazaki fragment is 1000-fold too slow to support an adequate rate of cycling and likely provides a backup mechanism to enable polymerase release when the other cycling signals are absent. Kinetic measurements indicate that addition of the last nucleotide to fill the gap is not the rate-limiting step for polymerase release and cycling. Modest (approximately 7 nt) strand displacement is observed after the gap between model Okazaki fragments is filled. To determine the identity of the protein that senses gap filling to modulate affinity of the replicase for the template, we performed photo-cross-linking experiments with highly reactive and non-chemoselective diazirines. Only the α subunit cross-linked, indicating that it serves as the sensor.  相似文献   

2.
Stable RNAs must fold into specific three-dimensional structures to be biologically active, yet many RNAs form metastable structures that compete with the native state. Our previous time-resolved footprinting experiments showed that Azoarcus group I ribozyme forms its tertiary structure rapidly (τ < 30 ms) without becoming significantly trapped in kinetic intermediates. Here, we use stopped-flow fluorescence spectroscopy to probe the global folding kinetics of a ribozyme containing 2-aminopurine in the loop of P9. The modified ribozyme was catalytically active and exhibited two equilibrium folding transitions centered at 0.3 and 1.6 mM Mg2+, consistent with previous results. Stopped-flow fluorescence revealed four kinetic folding transitions with observed rate constants of 100, 34, 1, and 0.1 s− 1 at 37 °C. From comparison with time-resolved Fe(II)-ethylenediaminetetraacetic acid footprinting of the modified ribozyme under the same conditions, these folding transitions were assigned to formation of the IC intermediate, tertiary folding and docking of the nicked P9 tetraloop, reorganization of the P3 pseudoknot, and refolding of nonnative conformers, respectively. The footprinting results show that 50-60% of the modified ribozyme folds in less than 30 ms, while the rest of the RNA population undergoes slow structural rearrangements that control the global folding rate. The results show how small perturbations to the structure of the RNA, such as a nick in P9, populate kinetic folding intermediates that are not observed in the natural ribozyme.  相似文献   

3.
A macromolecular nucleoprotein complex in retrovirus-infected cells, termed the preintegration complex, is responsible for the concerted integration of linear viral DNA genome into host chromosomes. Isolation of sufficient quantities of the cytoplasmic preintegration complexes for biochemical and biophysical analysis is difficult. We investigated the architecture of HIV-1 nucleoprotein complexes involved in the concerted integration pathway in vitro. HIV-1 integrase (IN) non-covalently juxtaposes two viral DNA termini forming the synaptic complex, a transient intermediate in the integration pathway, and shares properties associated with the preintegration complex. IN slowly processes two nucleotides from the 3′ OH ends and performs the concerted insertion of two viral DNA ends into target DNA. IN remains associated with the concerted integration product, termed the strand transfer complex. The synaptic complex and strand transfer complex can be isolated by native agarose gel electrophoresis. In-gel fluorescence resonance energy transfer measurements demonstrated that the energy transfer efficiencies between the juxtaposed Cy3 and Cy5 5′-end labeled viral DNA ends in the synaptic complex (0.68 ± 0.09) was significantly different from that observed in the strand transfer complex (0.07 ± 0.02). The calculated distances were 46 ± 3 Å and 83 ± 5 Å, respectively. DNaseI footprint analysis of the complexes revealed that IN protects U5 and U3 DNA sequences up to ∼ 32 bp from the end, suggesting two IN dimers were bound per terminus. Enhanced DNaseI cleavages were observed at nucleotide positions 6 and 9 from the terminus on U3 but not on U5, suggesting independent assembly events. Protein-protein cross-linking of IN within these complexes revealed the presence of dimers, tetramers, and a larger multimer (> 120 kDa). Our results suggest a new model where two IN dimers individually assemble on U3 and U5 ends before the non-covalent juxtaposition of two viral DNA ends, producing the synaptic complex.  相似文献   

4.
TRAPP complexes, which are large multimeric assemblies that function in membrane traffic, are guanine nucleotide exchange factors (GEFs) that activate the Rab GTPase Ypt1p. Here we measured rate and equilibrium constants that define the interaction of Ypt1p with guanine nucleotide (guanosine 5'-diphosphate and guanosine 5'-triphosphate/guanosine 5′-(β,γ-imido)triphosphate) and the core TRAPP subunits required for GEF activity. These parameters allowed us to identify the kinetic and thermodynamic bases by which TRAPP catalyzes nucleotide exchange from Ypt1p. Nucleotide dissociation from Ypt1p is slow (∼ 10− 4 s− 1) and accelerated > 1000-fold by TRAPP. Acceleration of nucleotide exchange by TRAPP occurs via a predominantly Mg2+-independent pathway. Thermodynamic linkage analysis indicates that TRAPP weakens nucleotide affinity by < 80-fold and vice versa, in contrast to most other characterized GEF systems that weaken nucleotide binding affinities by 4-6 orders of magnitude. The overall net changes in nucleotide binding affinities are small because TRAPP accelerates both nucleotide binding and dissociation from Ypt1p. Weak thermodynamic coupling allows TRAPP, Ypt1p, and nucleotide to exist as a stable ternary complex, analogous to strain-sensing cytoskeleton motors. These results illustrate a novel strategy of guanine nucleotide exchange by TRAPP that is particularly suited for a multifunctional GEF involved in membrane traffic.  相似文献   

5.
Integration of human immunodeficiency virus cDNA ends by integrase (IN) into host chromosomes involves a concerted integration mechanism. IN juxtaposes two DNA blunt ends to form the synaptic complex, which is the intermediate in the concerted integration pathway. The synaptic complex is inactivated by strand transfer inhibitors (STI) with IC50 values of ∼ 20 nM for inhibition of concerted integration. We detected a new nucleoprotein complex on a native agarose gel that was produced in the presence of > 200 nM STI, termed the IN-single DNA (ISD) complex. Two IN dimers appear to bind in a parallel fashion at the DNA terminus, producing an ∼ 32-bp DNase I protective footprint. In the presence of raltegravir (RAL), MK-2048, and L-841,411, IN incorporated ∼ 20-25% of the input blunt-ended DNA substrate into the stabilized ISD complex. Seven other STI also produced the ISD complex (≤ 5% of input DNA). The formation of the ISD complex was not dependent on 3′OH processing, and the DNA was predominantly blunt ended in the complex. The RAL-resistant IN mutant N155H weakly forms the ISD complex in the presence of RAL at ∼ 25% level of wild-type IN. In contrast, MK-2048 and L-841,411 produced ∼ 3-fold to 5-fold more ISD than RAL with N155H IN, which is susceptible to these two inhibitors. The results suggest that STI are slow-binding inhibitors and that the potency to form and stabilize the ISD complex is not always related to inhibition of concerted integration. Rather, the apparent binding and dissociation properties of each STI influenced the production of the ISD complex.  相似文献   

6.
7.
8.
An RNA kissing complex formed by the dimerization initiation site plays a critical role in the survival and infectivity of human immunodeficiency virus. Two dimerization initiation site kissing sequences, Mal and Lai, have been found in most human immunodeficiency virus 1 variants. Formation and stability of these RNA kissing complexes depend crucially on cationic conditions, particularly Mg2+. Using optical tweezers, we investigated the mechanical unfolding of single RNA molecules with either Mal-type (GUGCAC) or Lai-type (GCGCGC) kissing complexes under various ionic conditions. The force required to disrupt the kissing interaction of the two structures, the rip force, is sensitive to concentrations of KCl and MgCl2; addition of 3 mM MgCl2 to 100 mM KCl changes the rip force of Mal from 21 ± 4 to 46 ± 3 pN. From the rip force distribution, the kinetics of breaking the kissing interaction is calculated as a function of force and cation concentration. The two kissing complexes have distinct unfolding transition states, as shown by different values of ΔX, which is the distance from the folded structure to the unfolding transition state. The ΔX of Mal is ∼ 0.6 nm smaller than that of Lai, suggesting that fewer kissing base pairs are broken at the transition state of the former, consistent with observations that the Lai-type kissing complex is more stable and requires significantly more force to unfold than the Mal type. More importantly, neither K+ nor Mg2+ significantly changes the position of the transition state along the reaction coordinate. However, increasing concentrations of cations increase the kinetic barrier. We derived a cation-specific parameter, m, to describe how the height of the kinetic barrier depends on the concentration of cations. Our results suggest that Mg2+ greatly slows down the unfolding of the kissing complex but has moderate effects on the formation kinetics of the structure.  相似文献   

9.
The nonstructural protein 3 helicase (NS3h) of hepatitis C virus is a 3′-to-5′ superfamily 2 RNA and DNA helicase that is essential for the replication of hepatitis C virus. We have examined the kinetic mechanism of the translocation of NS3h along single-stranded nucleic acid with bases uridylate (rU), deoxyuridylate (dU), and deoxythymidylate (dT), and have found that the macroscopic rate of translocation is dependent on both the base moiety and the sugar moiety of the nucleic acid, with approximate macroscopic translocation rates of 3 nt s− 1 (oligo(dT)), 35 nt s− 1 (oligo(dU)), and 42 nt s− 1 (oligo(rU)), respectively. We found a strong correlation between the macroscopic translocation rates and the binding affinity of the translocating NS3h protein for the respective substrates such that weaker affinity corresponded to faster translocation. The values of K0.5 for NS3h translocation at a saturating ATP concentration are as follows: 3.3 ± 0.4 μM nucleotide (poly(dT)), 27 ± 2 μM nucleotide (poly(dU)), and 36 ± 2 μM nucleotide (poly(rU)). Furthermore, results of the isothermal titration of NS3h with these oligonucleotides suggest that differences in TΔS0 are the principal source of differences in the affinity of NS3h binding to these substrates. Interestingly, despite the differences in macroscopic translocation rates and binding affinities, the ATP coupling stoichiometries for NS3h translocation were identical for all three substrates (∼ 0.5 ATP molecule consumed per nucleotide translocated). This similar periodicity of ATP consumption implies a similar mechanism for NS3h translocation along RNA and DNA substrates.  相似文献   

10.
11.
Using cryo-electron microscopy, single particle image processing and three-dimensional reconstruction with icosahedral averaging, we have determined the three-dimensional solution structure of bacteriophage MS2 capsids reassembled from recombinant protein in the presence of short oligonucleotides. We have also significantly extended the resolution of the previously reported structure of the wild-type MS2 virion. The structures of recombinant MS2 capsids reveal clear density for bound RNA beneath the coat protein binding sites on the inner surface of the T = 3 MS2 capsid, and show that a short extension of the minimal assembly initiation sequence that promotes an increase in the efficiency of assembly, interacts with the protein capsid forming a network of bound RNA. The structure of the wild-type MS2 virion at ∼9 Å resolution reveals icosahedrally ordered density encompassing ∼90% of the single-stranded RNA genome. The genome in the wild-type virion is arranged as two concentric shells of density, connected along the 5-fold symmetry axes of the particle. This novel RNA fold provides new constraints for models of viral assembly.  相似文献   

12.
Escherichiacoli RecBCD is a bipolar DNA helicase possessing two motor subunits (RecB, a 3′-to-5′ translocase, and RecD, a 5′-to-3′ translocase) that is involved in the major pathway of recombinational repair. Previous studies indicated that the minimal kinetic mechanism needed to describe the ATP-dependent unwinding of blunt-ended DNA by RecBCD in vitro is a sequential n-step mechanism with two to three additional kinetic steps prior to initiating DNA unwinding. Since RecBCD can “melt out” ∼ 6 bp upon binding to the end of a blunt-ended DNA duplex in a Mg2+-dependent but ATP-independent reaction, we investigated the effects of noncomplementary single-stranded (ss) DNA tails [3′-(dT)6 and 5′-(dT)6 or 5′-(dT)10] on the mechanism of RecBCD and RecBC unwinding of duplex DNA using rapid kinetic methods. As with blunt-ended DNA, RecBCD unwinding of DNA possessing 3′-(dT)6 and 5′-(dT)6 noncomplementary ssDNA tails is well described by a sequential n-step mechanism with the same unwinding rate (mkU = 774 ± 16 bp s− 1) and kinetic step size (m = 3.3 ± 1.3 bp), yet two to three additional kinetic steps are still required prior to initiation of DNA unwinding (kC = 45 ± 2 s− 1). However, when the noncomplementary 5′ ssDNA tail is extended to 10 nt [5′-(dT)10 and 3′-(dT)6], the DNA end structure for which RecBCD displays optimal binding affinity, the additional kinetic steps are no longer needed, although a slightly slower unwinding rate (mkU = 538 ± 24 bp s− 1) is observed with a similar kinetic step size (m = 3.9 ± 0.5 bp). The RecBC DNA helicase (without the RecD subunit) does not initiate unwinding efficiently from a blunt DNA end. However, RecBC does initiate well from a DNA end possessing noncomplementary twin 5′-(dT)6 and 3′-(dT)6 tails, and unwinding can be described by a simple uniform n-step sequential scheme, without the need for the additional kC initiation steps, with a similar kinetic step size (m = 4.4 ± 1.7 bp) and unwinding rate (mkobs = 396 ± 15 bp s− 1). These results suggest that the additional kinetic steps with rate constant kC required for RecBCD to initiate unwinding of blunt-ended and twin (dT)6-tailed DNA reflect processes needed to engage the RecD motor with the 5′ ssDNA.  相似文献   

13.
The translocase of the outer mitochondrial membrane (TOM) complex is the main entry gate for proteins imported into mitochondria. We determined the structure of the native, unstained ∼ 550-kDa core-Tom20 complex from Saccharomycescerevisiae by cryo-electron microscopy at 18-Å resolution. The complex is triangular, measuring 145 Å on edge, and has near-3-fold symmetry. Its bulk is made up of three globular ∼ 50-Å domains. Three elliptical pores on the c-face merge into one central ∼ 70-Å cavity with a cage-like assembly on the opposite t-face. Nitrilotriacetic acid-gold labeling indicates that three Tom22 subunits in the TOM complex are located at the perimeter of the complex near the interface of the globular domains. We assign Tom22, which controls complex assembly, to three peripheral protrusions on the c-face, while the Tom20 subunit is tentatively assigned to the central protrusion on this surface. Based on our three-dimensional map, we propose a model of transient interactions and functional dynamics of the TOM assembly.  相似文献   

14.
A subclass of proteins with the SEA (sea urchin sperm protein, enterokinase, and agrin) domain fold exists as heterodimers generated by autoproteolytic cleavage within a characteristic G− 1S+ 1VVV sequence. Autoproteolysis occurs by a nucleophilic attack of the serine hydroxyl on the vicinal glycine carbonyl followed by an N → O acyl shift and hydrolysis of the resulting ester. The reaction has been suggested to be accelerated by the straining of the scissile peptide bond upon protein folding. In an accompanying article, we report the mechanism; in this article, we provide further key evidence and account for the energetics of coupled protein folding and autoproteolysis. Cleavage of the GPR116 domain and that of the MUC1 SEA domain occur with half-life (t½) values of 12 and 18 min, respectively, with lowering of the free energy of the activation barrier by ∼ 10 kcal mol− 1 compared with uncatalyzed hydrolysis. The free energies of unfolding of the GPR116 and MUC1 SEA domains were measured to ∼ 11 and ∼ 15 kcal mol− 1, respectively, but ∼ 7 kcal mol− 1 of conformational energy is partitioned as strain over the scissile peptide bond in the precursor to catalyze autoproteolysis by substrate destabilization. A straining energy of ∼ 7 kcal mol− 1 was measured by using both a pre-equilibrium model to analyze stability and cleavage kinetics data obtained with the GPR116 SEA domain destabilized by core mutations or urea addition, as well as the difference in thermodynamic stabilities of the MUC1 SEA precursor mutant S1098A (with a G− 1A+ 1VVV motif) and the wild-type protein. The results imply that cleavage by N → O acyl shift alone would proceed with a t½ of ∼ 2.3 years, which is too slow to be biochemically effective. A subsequent review of structural data on other self-cleaving proteins suggests that conformational strain of the scissile peptide bond may be a common mechanism of autoproteolysis.  相似文献   

15.
Adiponectin, a macromolecular complex similar to the members of the C1q and other collagenous homologues, elicits diverse biological functions, including anti-diabetes, anti-atherosclerosis, anti-inflammation and anti-tumor activities, which have been directly linked to the high molecular weight (HMW) oligomeric structures formed by multiples of adiponectin trimers. Here, we report the 3-D reconstructions of isolated full-length, recombinant murine C39A adiponectin trimer and hexamer of wild-type trimers (the major HMW form) determined by single-particle analysis of electron micrographs. The pleiomorphic ensemble of collagen-like stretches of the trimers leads to a dynamic structure of HMW that partition into two major classes, the fan-shaped (class I) and bouquet-shaped (class II). In both of these, while the N termini cluster into a compact ellipsoid-shaped (∼ 60 Å × 45 Å × 45 Å) volume, the collagenous domains assume a variety of arrangements. The domains are splayed by up to ∼ 90° in class I, can form a close-packed, up to ∼ 100 × 40 Å cylindrical assembly in class II, which can house about half of the 66 putative collagen-like sequence and the rest, tethered to the trimeric globular domains at the C terminus, are highly dynamic. As a result, the globular domains elaborate a variety of arrangements, covering an area of up to ∼ 4.9 × 105 Å2 and up to ∼ 320 Å apart, some of which were captured in reconstructions of class II. Our reconstructions suggest that the N-terminal structured domain, agreeing approximately with the expected volume for the octadecameric assembly of the terminal 27 amino acids, is crucial to the formation of the functionally active HMW. On the other hand, conformational flexibility of the trimers at the C terminus can allow the HMW to access and cluster disparate target ligands binding to the globular domains, which may be necessary to activate cellular signaling leading to the remarkable functional diversity of adiponectin.  相似文献   

16.
Several RNA-cleaving deoxyribozymes (DNAzymes) have been reported for efficient cleavage of purine-containing junctions, but none is able to efficiently cleave pyrimidine-pyrimidine (Pyr-Pyr) junctions. We hypothesize that a stronger Pyr-Pyr cleavage activity requires larger DNAzymes with complex structures that are difficult to isolate directly from a DNA library; one possible way to obtain such DNAzymes is to optimize DNA sequences with weak activities. To test this, we carried out an in vitro selection study to derive DNAzymes capable of cleaving an rC-T junction in a chimeric DNA/RNA substrate from DNA libraries constructed through chemical mutagenesis of five previous DNAzymes with a kobs of ∼ 0.001 min− 1 for the rC-T junction. After several rounds of selective amplification, DNAzyme descendants with a kobs of ∼ 0.1 min− 1 were obtained from a DNAzyme pool. The most efficient motif, denoted “CT10-3.29,” was found to have a catalytic core of ∼ 50 nt, larger than other known RNA-cleaving DNAzymes, and its secondary structure contains five short duplexes confined by a four-way junction. Several variants of CT10-3.29 exhibit a kobs of 0.3-1.4 min− 1 against the rC-T junction. CT10-3.29 also shows strong activity (kobs  > 0.1 min− 1) for rU-A and rU-T junctions, medium activity (> 0.01 min− 1) for rC-A and rA-T junctions, and weak activity (> 0.001 min− 1) for rA-A, rG-T, and rG-A junctions. Interestingly, a single-point mutation within the catalytic core of CT10-3.29 altered the pattern of junction specificity with a significantly decreased ability to cleave rC-T and rC-A junctions and a substantially increased ability to cleave rA-A, rA-T, rG-A, rG-T, rU-A, and rU-T junctions. This observation illustrates the intricacy and plasticity of this RNA-cleaving DNAzyme in dinucleotide junction selectivity. The current study shows that it is feasible to derive efficient DNAzymes for a difficult chemical task and reveals that DNAzymes require more complex structural solutions for such a task.  相似文献   

17.
Expression of the trypanosomal mitochondrial genome requires the insertion and deletion of uridylyl residues at specific sites in pre-mRNAs. RET2 terminal uridylyl transferase is an integral component of the RNA editing core complex (RECC) and is responsible for the guide-RNA-dependent U insertion reaction. By analyzing RNA-interference-based knock-in Trypanosoma brucei cell lines, purified editing complex, and individual protein, we have investigated RET2's association with the RECC. In addition, the U insertion activity exhibited by RET2 as an RECC subunit was compared with characteristics of the monomeric protein. We show that interaction of RET2 with RECC is accomplished via a protein-protein contact between its middle domain and a structural subunit, MP81. The recombinant RET2 catalyzes a faithful editing on gapped (precleaved) double-stranded RNA substrates, and this reaction requires an internal monophosphate group at the 5′ end of the mRNA 3′ cleavage fragment. However, RET2 processivity is limited to insertion of three Us. Incorporation into the RECC voids the internal phosphate requirement and allows filling of longer gaps similar to those observed in vivo. Remarkably, monomeric and RECC-embedded enzymes display a similar bimodal activity: the distributive insertion of a single uracil is followed by a processive extension limited by the number of guiding nucleotides. Based on the RNA substrate specificity of RET2 and the purine-rich nature of U insertion sites, we propose that the distributive + 1 insertion creates a substrate for the processive gap-filling reaction. Upon base-pairing of the + 1 extended 5′ cleavage fragment with a guiding nucleotide, this substrate is recognized by RET2 in a different mode compared to the product of the initial nucleolytic cleavage. Therefore, RET2 distinguishes base pairs in gapped RNA substrates which may constitute an additional checkpoint contributing to overall fidelity of the editing process.  相似文献   

18.
The folding mechanism and stability of dimeric formate dehydrogenase from Candida methylica was analysed by exposure to denaturing agents and to heat. Equilibrium denaturation data yielded a dissociation constant of about 10−13 M for assembly of the protein from unfolded chains and the kinetics of refolding and unfolding revealed that the overall process comprises two steps. In the first step a marginally stable folded monomeric state is formed at a rate (k1) of about 2 × 10−3 s−1 (by deduction k−1 is about10−4 s−1) and assembles into the active dimeric state with a bimolecular rate constant (k2) of about 2 × 104 M−1 s−1. The rate of dissociation of the dimeric state in physiological conditions is extremely slow (k−2 ∼ 3 × 10−7 s−1).  相似文献   

19.
20.
The rotavirus inner capsid particle, known as the “double-layered particle” (DLP), is the “payload” delivered into a cell in the process of viral infection. Its inner and outer protein layers, composed of viral protein (VP) 2 and VP6, respectively, package the 11 segments of the double-stranded RNA (dsRNA) of the viral genome, as well as about the same number of polymerase molecules (VP1) and capping-enzyme molecules (VP3). We have determined the crystal structure of the bovine rotavirus DLP. There is one full particle (outer diameter ∼ 700 Å) in the asymmetric unit of the P212121 unit cell of dimensions a = 740 Å, b = 1198 Å, and c = 1345 Å. A three-dimensional reconstruction from electron cryomicroscopy was used as a molecular replacement model for initial phase determination to about 18.5 Å resolution, and the 60-fold redundancy of icosahedral particle symmetry allowed phases to be extended stepwise to the limiting resolution of the data (3.8 Å). The structure of a VP6 trimer (determined previously by others) fits the outer layer density with very little adjustment. The T = 13 triangulation number of that layer implies that there are four and one-third VP6 trimers per icosahedral asymmetric unit. The inner layer has 120 copies of VP2 and thus 2 copies per icosahedral asymmetric unit, designated VP2A and VP2B. Residues 101-880 fold into a relatively thin principal domain, comma-like in outline, shaped such that only rather modest distortions (concentrated at two “subdomain” boundaries) allow VP2A and VP2B to form a uniform layer with essentially no gaps at the subunit boundaries, except for a modest pore along the 5-fold axis. The VP2 principal domain resembles those of the corresponding shells and homologous proteins in other dsRNA viruses: λ1 in orthoreoviruses and VP3 in orbiviruses. Residues 1-80 of VP2A and VP2B fold together with four other such pairs into a “5-fold hub” that projects into the DLP interior along the 5-fold axis; residues 81-100 link the 10 polypeptide chains emerging from a 5-fold hub to the N-termini of their corresponding principal domains, clustered into a decameric assembly unit. The 5-fold hub appears to have several distinct functions. One function is to recruit a copy of VP1 (or of a VP1-VP3 complex), potentially along with a segment of plus-strand RNA, as a decamer of VP2 assembles. The second function is to serve as a shaft around which can coil a segment of dsRNA. The third function is to guide nascent mRNA, synthesized in the DLP interior by VP1 and 5′-capped by the action of VP3, out through a 5-fold exit channel. We propose a model for rotavirus particle assembly, based on known requirements for virion formation, together with the structure of the DLP and that of VP1, determined earlier.  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