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 93 毫秒
1.
2.
3.
We have previously reported on the ubiquitylation and degradation of hepatitis C virus core protein. Here we demonstrate that proteasomal degradation of the core protein is mediated by two distinct mechanisms. One leads to polyubiquitylation, in which lysine residues in the N-terminal region are preferential ubiquitylation sites. The other is independent of the presence of ubiquitin. Gain- and loss-of-function analyses using lysineless mutants substantiate the hypothesis that the proteasome activator PA28γ, a binding partner of the core, is involved in the ubiquitin-independent degradation of the core protein. Our results suggest that turnover of this multifunctional viral protein can be tightly controlled via dual ubiquitin-dependent and -independent proteasomal pathways.Hepatitis C virus (HCV) core protein, whose amino acid sequence is highly conserved among different HCV strains, not only is involved in the formation of the HCV virion but also has a number of regulatory functions, including modulation of signaling pathways, cellular and viral gene expression, cell transformation, apoptosis, and lipid metabolism (reviewed in references 9 and 15). We have previously reported that the E6AP E3 ubiquitin (Ub) ligase binds to the core protein and plays an important role in polyubiquitylation and proteasomal degradation of the core protein (22). Another study from our group identified the proteasome activator PA28γ/REG-γ as an HCV core-binding partner, demonstrating degradation of the core protein via a PA28γ-dependent pathway (16, 17). In this work, we further investigated the molecular mechanisms underlying proteasomal degradation of the core protein and found that in addition to regulation by the Ub-mediated pathway, the turnover of the core protein is also regulated by PA28γ in a Ub-independent manner.Although ubiquitylation of substrates generally requires at least one Lys residue to serve as a Ub acceptor site (5), there is no consensus as to the specificity of the Lys targeted by Ub (4, 8). To determine the sites of Ub conjugation in the core protein, we used site-directed mutagenesis to replace individual Lys residues or clusters of Lys residues with Arg residues in the N-terminal 152 amino acids (aa) of the core (C152), within which is contained all seven Lys residues (Fig. (Fig.1A).1A). Plasmids expressing a variety of mutated core proteins were generated by PCR and inserted into the pCAGGS (18). Each core-expressing construct was transfected into human embryonic kidney 293T cells along with the pMT107 (25) encoding a Ub moiety tagged with six His residues (His6). Transfected cells were treated with the proteasome inhibitor MG132 for 14 h to maximize the level of Ub-conjugated core intermediates by blocking the proteasome pathway and were harvested 48 h posttransfection. His6-tagged proteins were purified from the extracts by Ni2+-chelation chromatography. Eluted protein and whole lysates of transfected cells before purification were analyzed by Western blotting using anticore antibodies (Fig. (Fig.1B).1B). Mutations replacing one or two Lys residues with Arg in the core protein did not affect the efficiency of ubiquitylation: detection of multiple Ub-conjugated core intermediates was observed in the mutant core proteins comparable to the results seen with the wild-type core protein as previously reported (23). In contrast, a substitution of four N-terminal Lys residues (C152K6-23R) caused a significant reduction in ubiquitylation (Fig. (Fig.1B,1B, lane 9). Multiple Ub-conjugated core intermediates were not detected in the Lys-less mutant (C152KR), in which all seven Lys residues were replaced with Arg (Fig. (Fig.1B,1B, lane 11). These results suggest that there is not a particular Lys residue in the core protein to act as the Ub acceptor but that more than one Lys located in its N-terminal region can serve as the preferential ubiquitylation site. In rare cases, Ub is known to be conjugated to the N terminus of proteins; however, these results indicate that this does not occur within the core protein.Open in a separate windowFIG. 1.In vivo ubiquitylation of HCV core protein. (A) The HCV core protein (N-terminal 152 aa) is represented on the top. The positions of the amino acid residues of the core protein are indicated above the bold lines. The positions of the seven Lys residues in the core are marked by vertical ticks. Substitution of Lys with Arg (R) is schematically depicted. (B) Detection of ubiquitylated forms of the core proteins. The transfected cells with core expression plasmids and pMT107 were treated with the proteasome inhibitor MG132 and harvested 48 h after transfection. His6-tagged proteins were purified and subsequently analyzed by Western blot analysis using anticore antibody (upper panel). Core proteins conjugated to a number of His6-Ub are denoted with asterisks. Whole lysates of transfected cells before purification were also analyzed (lower panel). Lanes 1 to 11, C152 to C152KR, as indicated for panel A. Lane 12; empty vector.To investigate how polyubiquitylation correlates with proteasome degradation of the core protein, we performed kinetic analysis of the wild-type and mutated core proteins by use of the Ub protein reference (UPR) technique, which can compensate for data scatter of sample-to-sample variations such as levels of expression (10, 24). Fusion proteins expressed from UPR-based constructs (Fig. (Fig.2A)2A) were cotranslationally cleaved by deubiquitylating enzymes, thereby generating equimolar quantities of the core proteins and the reference protein, dihydrofolate reductase-hemagglutinin (DHFR-HA) tag-modified Ub, in which the Lys at aa 48 was replaced by Arg to prevent its polyubiquitylation (UbR48). After 24 h of transfection with UPR constructs, cells were treated with cycloheximide and the amounts of core proteins and DHFR-HA-UbR48 at the indicated time points were determined by Western blot analysis using anticore and anti-HA antibodies. The mature form of the core protein, aa 1 to 173 (C173) (13, 20), and C152 were degraded with first-order kinetics (Fig. 2B and D). MG132 completely blocked the degradation of C173 and C152 (Fig. (Fig.2B),2B), and C152K6-23R and C152KR were markedly stabilized (Fig. (Fig.2C).2C). The half-lives of C173 and C152 were calculated to be 5 to 6 h, whereas those of C152K6-23R and C152KR were calculated to be 22 to 24 h (Fig. (Fig.2D),2D), confirming that the Ub plays an important role in regulating degradation of the core protein. Nevertheless, these results also suggest possible involvement of the Ub-independent pathway in the turnover of the core protein, as C152KR is more destabilized than the reference protein (Fig. (Fig.2C2C and and2D2D).Open in a separate windowFIG. 2.Kinetic analysis of degradation of HCV core proteins. (A) The fusion constructs used in the UPR technique. Open boxes indicate the DHFR sequence, which is extended at the C terminus by a sequence containing the HA epitope (hatched boxes). UbR48 moieties bearing the Lys-Arg substitution at aa 48 are represented by open ellipses. Bold lines indicate the regions of the core protein. The amino acid positions of the core protein are indicated above the bold lines. The arrows indicate the sites of in vivo cleavage by deubiquitylating enzymes. (B and C) Turnover of the core proteins. After a 24-h transfection with each UPR construct, cells were treated with 50 μg of cycloheximide/ml in the presence or absence of 10 μM MG132 for the different time periods indicated. Cells were lysed at the different time points indicated, followed by evaluation via sodium dodecyl sulfate-polyacrylamide gel electrophoresis and Western blot analysis using antibodies against the core protein and HA. (D) Quantitation of the data shown in panels B and C. At each time point, the ratio of band intensity of the core protein relative to the reference DHFR-HA-UbR48 was determined by densitometry and is plotted as a percentage of the ratio at time zero.We have shown that PA28γ specifically binds to the core protein and is involved in its degradation (16, 17). Recent studies demonstrated that PA28γ is responsible for Ub-independent degradation of the steroid receptor coactivator SRC-3 and cell cycle inhibitors such as p21 (3, 11, 12). Thus, we next investigated the possibility of PA28γ involvement in the degradation of either C152KR or C152. Since C152KR carries two amino acid substitutions in the PA28γ-binding region (aa 44 to 71) (17), we tested the influence of the mutations of C152KR on the interaction with PA28γ by use of a coimmunoprecipitation assay. When Flag-tagged PA28γ (F-PA28γ) was expressed in cells along with C152 or C152KR, F-PA28γ precipitated along with both C152 and C152KR, indicating that PA28γ interacts with both core proteins (Fig. (Fig.3A).3A). Figure Figure3B3B reveals the effect of exogenous expression of F-PA28γ on the steady-state levels of C152 and C152KR. Consistent with previous data (17), the expression level of C152 was decreased to a nearly undetectable level in the presence of PA28γ (Fig. (Fig.3B,3B, lanes 1 and 3). Interestingly, exogenous expression of PA28γ led to a marked reduction in the amount of C152KR expressed (Fig. (Fig.3B,3B, lanes 5 and 7). Treatment with MG132 increased the steady-state level of the C152KR in the presence of F-PA28γ as well as the level of C152 (Fig. (Fig.3B,3B, lanes 4 and 8).Open in a separate windowFIG. 3.PA28γ-dependent degradation of the core protein. (A) Interaction of the core protein with PA28γ. Cells were cotransfected with the wild-type (C152) or Lys-less (C152KR) core expression plasmid in the presence of a Flag-PA28γ (F-PA28γ) expression plasmid or an empty vector. The transfected cells were treated with MG132. After 48 h, the cell lysates were immunoprecipitated with anti-Flag antibody and visualized by Western blotting with anticore antibodies. Western blot analysis of whole cell lysates was also performed. (B) Degradation of the wild-type and Lys-less core proteins via the PA28γ-dependent pathway. Cells were transfected with the UPR construct with or without F-PA28γ. In some cases, cells were treated with 10 μM MG132 for 14 h before harvesting. Western blot analysis was performed using anticore, anti-HA, and anti-Flag antibodies. (C) After 24 h of transfection with UPR-C152KR and UPR-C191KR with or without F-PA28γ (an empty vector), cells were treated with 50 μg of cycloheximide/ml for different time periods as indicated (chase time). Western blot analysis was performed using anticore and anti-HA antibodies. The precursor core protein and the core that was processed, presumably by signal peptide peptidase, are denoted by open and closed triangles, respectively.We further investigated whether PA28γ affects the turnover of Lys-less core protein through time course experiments. C152KR was rapidly destabilized and almost completely degraded in a 3-h chase experiment using cells overexpressing F-PA28γ (Fig. (Fig.3C,3C, left panels). A similar result was obtained using an analogous Lys-less mutant of the full-length core protein C191KR (Fig. (Fig.3C,3C, right panels), thus demonstrating that the Lys-less core protein undergoes proteasomal degradation in a PA28γ-dependent manner. These results suggest that PA28γ may play a role in accelerating the turnover of the HCV core protein that is independent of ubiquitylation.Finally, we examined gain- and loss-of-function of PA28γ with respect to degradation of full-length wild-type (C191) and mutated (C191KR) core proteins in human hepatoma Huh-7 cells. As expected, exogenous expression of PA28γ or E6AP caused a decrease in the C191 steady-state levels (Fig. (Fig.4A).4A). In contrast, the C191KR level was decreased with expression of PA28γ but not of E6AP. We further used RNA interference to inhibit expression of PA28γ or E6AP. An increase in the abundance of C191KR was observed with PA28γ small interfering RNA (siRNA) but not with E6AP siRNA (Fig. (Fig.4B).4B). An increase in the C191 level caused by the activity of siRNA against PA28γ or E6AP was confirmed as well.Open in a separate windowFIG. 4.Ub-dependent and Ub-independent degradation of the full-length core protein in hepatic cells. (A) Huh-7 cells were cotransfected with plasmids for the full-length core protein (C191) or its Lys-less mutant (C191KR) in the presence of F-PA28γ or HA-tagged-E6AP expression plasmid (HA-E6AP). After 48 h, cells were lysed and Western blot analysis was performed using anticore, anti-HA, anti-Flag, or anti-GAPDH. (B) Huh-7 cells were cotransfected with core expression plasmids along with siRNA against PA28γ or E6AP or with negative control siRNA. Cells were harvested 72 h after transfection and subjected to Western blot analysis.Taking these results together, we conclude that turnover of the core protein is regulated by both Ub-dependent and Ub-independent pathways and that PA28γ is possibly involved in Ub-independent proteasomal degradation of the core protein. PA28 is known to specifically bind and activate the 20S proteasome (19). Thus, PA28γ may function by facilitating the delivery of the core protein to the proteasome in a Ub-independent manner.Accumulating evidence suggests the existence of proteasome-dependent but Ub-independent pathways for protein degradation, and several important molecules, such as p53, p73, Rb, SRC-3, and the hepatitis B virus X protein, have two distinct degradation pathways that function in a Ub-dependent and Ub-independent manner (1, 2, 6, 7, 14, 21, 27). Recently, critical roles for PA28γ in the Ub-independent pathway have been demonstrated; SRC-3 and p21 can be recognized by the 20S proteasome independently of ubiquitylation through their interaction with PA28γ (3, 11, 12). It has also been reported that phosphorylation-dependent ubiquitylation mediated by GSK3 and SCF is important for SRC-3 turnover (26). Nevertheless, the precise mechanisms underlying turnover of most of the proteasome substrates that are regulated in both Ub-dependent and Ub-independent manners are not well understood. To our knowledge, the HCV core protein is the first viral protein studied that has led to identification of key cellular factors responsible for proteasomal degradation via dual distinct mechanisms. Although the question remains whether there is a physiological significance of the Ub-dependent and Ub-independent degradation of the core protein, it is reasonable to consider that tight control over cellular levels of the core protein, which is multifunctional and essential for viral replication, maturation, and pathogenesis, may play an important role in representing the potential for its functional activity.  相似文献   

4.
5.
6.
7.
8.
Our previous structural studies on intact, infectious murine norovirus 1 (MNV-1) virions demonstrated that the receptor binding protruding (P) domains are lifted off the inner shell of the virus. Here, the three-dimensional (3D) reconstructions of recombinant rabbit hemorrhagic disease virus (rRHDV) virus-like particles (VLPs) and intact MNV-1 were determined to ∼8-Å resolution. rRHDV also has a raised P domain, and therefore, this conformation is independent of infectivity and genus. The atomic structure of the MNV-1 P domain was used to interpret the MNV-1 reconstruction. Connections between the P and shell domains and between the floating P domains were modeled. This observed P-domain flexibility likely facilitates virus-host receptor interactions.Murine norovirus 1 (MNV-1) (3, 14, 15) and rabbit hemorrhagic disease virus (RHDV) are members of the genera Norovirus and Lagovirus of the family Caliciviridae that offer a comparison to recombinant human norovirus (rNV) virus-like particles (VLPs) for assessing the structures and roles of domains within the capsid proteins of this family of viruses. Calicivirus particles contain 180 copies of the 56- to 76-kDa major capsid protein (Orf2), which is comprised of the internal/buried N terminus (N), shell (S), and protruding (P) domains (9, 10). The S domain, an eight-stranded β-barrel, forms an ∼300-Å contiguous shell around the RNA genome. A flexible hinge connects the shell to a “protruding” (P) domain at the C-terminal half of the capsid protein, which can be further divided into a globular head region (P2) and a stem region (P1) that connects the shell domain to P2. The accompanying article (13) describes the determination of the structure of the P domain of MNV-1 to a resolution of 2.0 Å.We recently determined the cryo-transmission electron microscopy (TEM) structure of MNV-1 to ∼12-Å resolution (4) and found that, compared to rNV VLPs (10) and San Miguel sea lion virus (SMSV) (1, 2), the protruding domains are rotated by ∼40° in a clockwise fashion and lifted up by ∼16 Å. To better understand the unusual conformation of MNV-1 and whether it is unique to this particular member of the calicivirus family, the ∼8-Å cryo-TEM structures of infectious MNV-1 and the VLPs of RHDV were determined.MNV-1 was produced as previously described (4). Three liters of cell culture yielded 0.5 to 1.0 mg of purified virus with a particle/PFU ratio of less than 100. Baculovirus expression and purification of recombinant RHDV (rRHDV) VLPs were performed as previously described (8). Cryo-electron microscopy (EM) data were collected at the National Resource for Automated Molecular Microscopy (NRAMM) facility in San Diego, CA (4). Images were collected at a nominal magnification of ×50,000 at a pixel size of 0.1547 nm at the specimen level using Leginon software (12) and processed with Appion software (5). The contrast transfer function for each set of particles from each image was estimated and corrected using ACE2 (a variation of ACE [7]). Particle images were automatically selected (11). The final stacks of particle images contained 20,425 MNV virions and 7,856 rRHDV VLPs, and EMAN 3D (6) was used for the reconstructions. Resolutions were estimated by Fourier shell correlations (FSC) of the three-dimensional (3D) reconstructions and application of a cutoff of 0.5. An amplitude correction of the final electron density was performed using GroEL small-angle X-ray scattering (SAXS) data.3D reconstructions of MNV-1 and rRHDV were calculated to resolutions of 8 Å and 8.1 Å, respectively (Fig. (Fig.1).1). The P domains of rNV VLPs rest directly on top of the shell domain (10) (Fig. (Fig.1A).1A). In contrast, the P domains of MNV-1 are lifted and rotated above the shell of the capsid (4) (Fig. (Fig.1B).1B). At this higher resolution, there was a clear connection between the P1 domain and the shell domain in all three capsid subunits (Fig. (Fig.1B,1B, arrow A). Unlike the smooth protruding domains of rNV, MNV-1 has two clear “horns” (arrow B), not dissimilar to those observed for the sapoviruses (1, 2). There also are islands of density in the interior of the shell, directly beneath the 5-fold axes, that may represent ordered regions of RNA.Open in a separate windowFIG. 1.Stereo diagrams (left) and thin sections (right), with radius coloring, of rNV (A), MNV-1 (B), and an rRHDV VLP (C). For rNV, the atomic coordinates (10) were used. In MNV, arrow A indicates the thin connector between the P1 and S domains. Arrow B denotes the horns found at the tips of the P2 domains. Arrow C denotes the large gap between the P1 and S domains in the rRHDV VLP. Arrow D denotes the false connectivity in rRHDV VLPs between the P1 domain and the S domain near the 5-fold axes.As with MNV-1, there is a marked gap between the P and S domains in the rRHDV VLP (Fig. (Fig.1C,1C, arrow C). This gap is not as pronounced as in MNV-1 because the P domains are not rotated as in MNV-1. In this electron density map, the A/B dimers appear to be touching the shell domain near the 5-fold axes. This contact difference between the A/B dimers and the C/C dimers could be the reason why the tops of the C/C dimers appear to be markedly disordered compared to the A/B dimers in rRHDV and the C/C dimers in MNV-1.Shown in Fig. Fig.22 is the fitting of the atomic structures of the MNV-1 P domain (13) and the rNV S domains into the MNV-1 3D reconstruction electron density. The horns (arrow A, loops A′-B′ and E′-F′) observed at the tips of the P domain match exceedingly well with the electron density. As discussed in the accompanying publication (13), the A′-B′ and E′-F′ loops displayed two discrete conformations, a closed structure, where the two loops were tightly associated, and an open structure, where the loops were splayed apart. The horns of the closed conformation fit better into the reconstruction, as the E′-F′ loop in the open form jutted out of the density at the base of the horns. The unmodified density in the lower panel of Fig. Fig.22 shows fine features in the shell domain and a very clear connection between the shell and P1 domains. The connections between the P1 and S domains were of sufficient quality to build a basic backbone model by uncoiling the linker region (arrow B). The P domain in the unfiltered 3D reconstruction was far less ordered than the S domain (Fig. (Fig.2).2). This was likely due to movement of the entire P domain with respect to the shell.Open in a separate windowFIG. 2.Fitting of the MNV-1 P domain and the rNV shell domain into the MNV-1 electron density. A, B, and C subunits are represented by blue, green, and red, respectively. The electron density is shown in transparent gray. The top panel is the 8.0-Å-resolution 3D reconstruction modified using a low-pass filter. The bottom panel is the reconstruction without modification. The horns on the tops of the P domains are denoted by arrow A. Arrow B denotes the connection between the S and P domains.Using the structure of rNV VLP P domains for modeling, the rRHDV P domains are lifted off the surface of the shell, but not rotated as with MNV-1. This places the bottom edge of the A subunit P1 domain near the S domain at the 5-fold axes. The P-domain dimers of rNV and rRHDV have a more “arch-like” shape than MNV-1. Unlike in MNV-1, the electron densities of the C/C dimers in rRHDV are far more diffuse than those of the A/B dimers (Fig. (Fig.3B)3B) and the connector between the S and P1 domains is not clear. During fitting, the connector region was not as extended as with MNV-1. This may afford greater flexibility, leading to more diffuse electron density.Open in a separate windowFIG. 3.Fitting of the rNV atomic structure into the rRHDV VLP electron density. The upper stereo image shows the 8.1-Å-resolution 3D reconstruction after modification by a low-pass filter. Below is the same reconstruction prior to density modification.When the atomic models for the MNV-1 P domains (13) were placed into the cryo-TEM electron density (Fig. (Fig.4),4), the C termini extended deep into the cores of adjacent P domains. Possible connections not accounted for by the P-domain structures were also observed in the electron density between the P domains. A bulge between the P1 and P2 domains in the 3D reconstruction indicated a possible interaction between the C termini and the adjacent P domains. These same interactions were observed in the crystal lattice. This highly mobile C terminus may be a flexible tether between the P domains in the intact virion.Open in a separate windowFIG. 4.Possible carboxyl-terminus interactions between the P domains of MNV-1. (A) Stereo image of MNV-1 calculated to 12-Å resolution with (red) and without (yellow) the last 10 residues of the P domain. (B) The calculated MNV-1 density with the carboxyl terminus removed (yellow) overlaid onto the 3D reconstruction of MNV-1 (blue). Note the strands of difference density that roughly correspond to the C terminus in panel A. (C) The C-terminus interactions observed in the structure of the MNV-1 P domains. Shown in blue and green are ribbon diagrams of an A/B P-domain dimer. In mauve is a surface rendering of the C terminus from a crystallographically related dimer. (D) Surface rendering of the final MNV-1 model with possible interactions between the P domains in MNV-1. The carboxyl termini of the A subunits (blue) interact with the counterclockwise-related B subunits around the 5-fold axes (white arrows). Around the 3-fold (quasi-6-fold) axes, the C subunits interact with the A subunits and the B subunits interact with the C subunits (orange arrows).It is absolutely clear that the hinge region between the S and P domains affords a remarkable degree of flexibility in the P domains that is not genus specific or related to differences between rVLPs and authentic virions. The simplest explanation for the role of this transition is that it gives the P domains flexibility that may be used to optimize interactions with cell receptors during attachment and entry. In this way, the P domains can increase their avidity for the cell surface by being more facile in adapting to the presentation of cellular recognition motifs.  相似文献   

9.
10.
11.
12.
Streptomyces development was analyzed under conditions resembling those in soil. The mycelial growth rate was much lower than that in standard laboratory cultures, and the life span of the previously named first compartmentalized mycelium was remarkably increased.Streptomycetes are gram-positive, mycelium-forming, soil bacteria that play an important role in mineralization processes in nature and are abundant producers of secondary metabolites. Since the discovery of the ability of these microorganisms to produce clinically useful antibiotics (2, 15), they have received tremendous scientific attention (12). Furthermore, its remarkably complex developmental features make Streptomyces an interesting subject to study. Our research group has extended our knowledge about the developmental cycle of streptomycetes, describing new aspects, such as the existence of young, fully compartmentalized mycelia (5-7). Laboratory culture conditions (dense inocula, rich culture media, and relatively elevated temperatures [28 to 30°C]) result in high growth rates and an orderly-death process affecting these mycelia (first death round), which is observed at early time points (5, 7).In this work, we analyzed Streptomyces development under conditions resembling those found in nature. Single colonies and soil cultures of Streptomyces antibioticus ATCC 11891 and Streptomyces coelicolor M145 were used for this analysis. For single-colony studies, suitable dilutions of spores of these species were prepared before inoculation of plates containing GYM medium (glucose, yeast extract, malt extract) (11) or GAE medium (glucose, asparagine, yeast extract) (10). Approximately 20 colonies per plate were obtained. Soil cultures were grown in petri dishes with autoclaved oak forest soil (11.5 g per plate). Plates were inoculated directly with 5 ml of a spore suspension (1.5 × 107 viable spores ml−1; two independent cultures for each species). Coverslips were inserted into the soil at an angle, and the plates were incubated at 30°C. To maintain a humid environment and facilitate spore germination, the cultures were irrigated with 3 ml of sterile liquid GAE medium each week.The development of S. coelicolor M145 single colonies growing on GYM medium is shown in Fig. Fig.1.1. Samples were collected and examined by confocal microscopy after different incubation times, as previously described (5, 6). After spore germination, a viable mycelium develops, forming clumps which progressively extend along the horizontal (Fig. 1a and b) and vertical (Fig. 1c and d) axes of a plate. This mycelium is fully compartmentalized and corresponds to the first compartmentalized hyphae previously described for confluent surface cultures (Fig. 1e, f, and j) (see below) (5); 36 h later, death occurs, affecting the compartmentalized hyphae (Fig. 1e and f) in the center of the colony (Fig. (Fig.1g)1g) and in the mycelial layers below the mycelial surface (Fig. 1d and k). This death causes the characteristic appearance of the variegated first mycelium, in which alternating live and dead segments are observed (Fig. 1f and j) (5). The live segments show a decrease in fluorescence, like the decrease in fluorescence that occurs in solid confluent cultures (Fig. (Fig.11 h and i) (5, 9). As the cycle proceeds, the intensity of the fluorescence in these segments returns, and the segments begin to enlarge asynchronously to form a new, multinucleated mycelium, consisting of islands or sectors on the colony surfaces (Fig. 1m to o). Finally, death of the deeper layers of the colony (Fig. (Fig.1q)1q) and sporulation (Fig. (Fig.1r)1r) take place. Interestingly, some of the spores formed germinate (Fig. (Fig.1s),1s), giving rise to a new round of mycelial growth, cell death, and sporulation. This process is repeated several times, and typical, morphologically heterogeneous Streptomyces colonies grow (not shown). The same process was observed for S. antibioticus ATCC 11891, with minor differences mainly in the developmental time (not shown).Open in a separate windowFIG. 1.Confocal laser scanning fluorescence microscopy analysis of the development-related cell death of S. coelicolor M145 in surface cultures containing single colonies. Developmental culture times (in hours) are indicated. The images in panels l and n were obtained in differential interference contrast mode and correspond to the same fields as in panels k and m, respectively. The others are culture sections stained with SYTO 9 and propidium iodide. Panels c, d, k, l, p, and q are cross sections; the other images are longitudinal sections (see the methods). Panels h and i are images of the same field taken with different laser intensities, showing low-fluorescence viable hyphae in the center of the colonies that develop into a multinucleated mycelium. The arrows in panels e and s indicate septa (e) and germinated spores (s). See the text for details.Figure Figure22 shows the different types of mycelia present in S. coelicolor cultures under the conditions described above, depending on the compartmentalization status. Hyphae were treated with different fluorescent stains (SYTO 9 plus propidium iodide for nucleic acids, CellMask plus FM4-64 for cell membranes, and wheat germ agglutinin [WGA] for cell walls). Samples were processed as previously described (5). The young initial mycelia are fully compartmentalized and have membranous septa (Fig. 2b to c) with little associated cell wall material that is barely visible with WGA (Fig. (Fig.2d).2d). In contrast, the second mycelium is a multinucleated structure with fewer membrane-cell wall septa (Fig. 2e to h). At the end of the developmental cycle, multinucleated hyphae begin to undergo the segmentation which precedes the formation of spore chains (Fig. 2i to m). Similar results were obtained for S. antibioticus (not shown), but there were some differences in the numbers of spores formed. Samples of young and late mycelia were freeze-substituted using the methodology described by Porta and Lopez-Iglesias (13) and were examined with a transmission electron microscope (Fig. 2n and o). The septal structure of the first mycelium (Fig. (Fig.2n)2n) lacks the complexity of the septal structure in the second mycelium, in which a membrane with a thick cell wall is clearly visible (Fig. (Fig.2o).2o). These data coincide with those previously described for solid confluent cultures (4).Open in a separate windowFIG. 2.Analysis of S. coelicolor hyphal compartmentalization with several fluorescent indicators (single colonies). Developmental culture times (in hours) are indicated. (a, e, and i) Mycelium stained with SYTO 9 and propidium iodide (viability). (b, f, and j) Hyphae stained with Cell Mask (a membrane stain). (c, g, and l) Hyphae stained with FM 4-64 (a membrane stain). (d, h, and m) Hyphae stained with WGA (cell wall stain). Septa in all the images in panels a to j, l, and m are indicated by arrows. (k) Image of the same field as panel j obtained in differential interference contrast mode. (n and o) Transmission electron micrographs of S. coelicolor hyphae at different developmental phases. The first-mycelium septa (n) are comprised of two membranes separated by a thin cell wall; in contrast, second-mycelium septa have thick cell walls (o). See the text for details. IP, propidium iodide.The main features of S. coelicolor growing in soils are shown in Fig. Fig.3.3. Under these conditions, spore germination is a very slow, nonsynchronous process that commences at about 7 days (Fig. 3c and d) and lasts for at least 21 days (Fig. 3i to l), peaking at around 14 days (Fig. 3e to h). Mycelium does not clump to form dense pellets, as it does in colonies; instead, it remains in the first-compartmentalized-mycelium phase during the time analyzed. Like the membrane septa in single colonies, the membrane septa of the hyphae are stained with FM4-64 (Fig. 3j and k), although only some of them are associated with thick cell walls (WGA staining) (Fig. (Fig.3l).3l). Similar results were obtained for S. antibioticus cultures (not shown).Open in a separate windowFIG. 3.Confocal laser scanning fluorescence microscopy analysis of the development-related cell death and hyphal compartmentalization of S. coelicolor M145 growing in soil. Developmental culture times (in days) are indicated. The images in panels b, f, and h were obtained in differential interference contrast mode and correspond to the same fields as the images in panels a, e, and g, respectively. The dark zone in panel h corresponds to a particle of soil containing hyphae. (a, c, d, e, g, i, j, and k) Hyphae stained with SYTO 9, propidium iodide (viability stain), and FM4-64 (membrane stain) simultaneously. (i) SYTO 9 and propidium iodide staining. (j) FM4-64 staining. The image in panel k is an overlay of the images in panels i and j and illustrates that first-mycelium membranous septa are not always apparent when they are stained with nucleic acid stains (SYTO 9 and propidium iodide). (l) Hyphae stained with WGA (cell wall stain), showing the few septa with thick cell walls present in the cells. Septa are indicated by arrows. IP, propidium iodide.In previous work (8), we have shown that the mycelium currently called the substrate mycelium corresponds to the early second multinucleated mycelium, according to our nomenclature, which still lacks the hydrophobic layers characteristic of the aerial mycelium. The aerial mycelium therefore corresponds to the late second mycelium which has acquired hydrophobic covers. This multinucleated mycelium as a whole should be considered the reproductive structure, since it is destined to sporulate (Fig. (Fig.4)4) (8). The time course of lysine 6-aminotransferase activity during cephamycin C biosynthesis has been analyzed by other workers using isolated colonies of Streptomyces clavuligerus and confocal microscopy with green fluorescent protein as a reporter (4). A complex medium and a temperature of 29°C were used, conditions which can be considered similar to the conditions used in our work. Interestingly, expression did not occur during the development of the early mycelium and was observed in the mycelium only after 80 h of growth. This suggests that the second mycelium is the antibiotic-producing mycelium, a hypothesis previously confirmed using submerged-growth cultures of S. coelicolor (9).Open in a separate windowFIG. 4.Cell cycle features of Streptomyces growing under natural conditions. Mycelial structures (MI, first mycelium; MII, second mycelium) and cell death are indicated. The postulated vegetative and reproductive phases are also indicated (see text).The significance of the first compartmentalized mycelium has been obscured by its short life span under typical laboratory culture conditions (5, 6, 8). In previous work (3, 7), we postulated that this structure is the vegetative phase of the bacterium, an hypothesis that has been recently corroborated by proteomic analysis (data not shown). Death in confluent cultures begins shortly after germination (4 h) and continues asynchronously for 15 h. The second multinucleated mycelium emerges after this early programmed cell death and is the predominant structure under these conditions. In contrast, as our results here show, the first mycelium lives for a long time in isolated colonies and soil cultures. As suggested in our previous work (5, 6, 8), if we assume that the compartmentalized mycelium is the Streptomyces vegetative growth phase, then this phase is the predominant phase in individual colonies (where it remains for at least 36 h), soils (21 days), and submerged cultures (around 20 h) (9). The differences in the life span of the vegetative phase could be attributable to the extremely high cell densities attained under ordinary laboratory culture conditions, which provoke massive differentiation and sporulation (5-7, 8).But just exactly what are “natural conditions”? Some authors have developed soil cultures of Streptomyces to study survival (16, 17), genetic transfer (14, 17-19), phage-bacterium interactions (3), and antibiotic production (1). Most of these studies were carried out using amended soils (supplemented with chitin and starch), conditions under which growth and sporulation were observed during the first few days (1, 17). These conditions, in fact, might resemble environments that are particularly rich in organic matter where Streptomyces could conceivably develop. However, natural growth conditions imply discontinuous growth and limited colony development (20, 21). To mimic such conditions, we chose relatively poor but more balanced carbon-nitrogen soil cultures (GAE medium-amended soil) and less dense spore inocula, conditions that allow longer mycelium growth times. Other conditions assayed, such as those obtained by irrigating the soil with water alone, did not result in spore germination and mycelial growth (not shown). We were unable to detect death, the second multinucleated mycelium described above, or sporulation, even after 1 month of incubation at 30°C. It is clear that in nature, cell death and sporulation must take place at the end of the long vegetative phase (1, 17) when the imbalance of nutrients results in bacterial differentiation.In summary, the developmental kinetics of Streptomyces under conditions resembling conditions in nature differs substantially from the developmental kinetics observed in ordinary laboratory cultures, a fact that should be born in mind when the significance of development-associated phenomena is analyzed.  相似文献   

13.
Vector insect cells infected with Rice gall dwarf virus, a member of the family Reoviridae, contained the virus-associated microtubules adjacent to the viroplasms, as revealed by transmission electron, electron tomographic, and confocal microscopy. The viroplasms, putative sites of viral replication, contained the nonstructural viral proteins Pns7 and Pns12, as well as core protein P5, of the virus. Microtubule-depolymerizing drugs suppressed the association of viral particles with microtubules and prevented the release of viruses from cells without significantly affecting viral multiplication. Thus, microtubules appear to mediate viral transport within and release of viruses from infected vector cells.Rice gall dwarf virus (RGDV), Rice dwarf virus (RDV), and Wound tumor virus, members of the genus Phytoreovirus in the family Reoviridae, multiply both in plants and in invertebrate insect vectors. Each virus exists as icosahedral particles of approximately 65 to 70 nm in diameter, with two concentric layers (shells) of proteins that enclose a core (1, 13). The viral genome of RGDV consists of 12 segmented double-stranded RNAs that encode six structural (P1, P2, P3, P5, P6, and P8) and six nonstructural (Pns4, Pns7, Pns9, Pns10, Pns11, and Pns12) proteins (reference 21 and references therein). The core capsid is composed of P3, the major protein, which encloses P1, P5, and P6 (12). The outer layer consists of two proteins, namely, P2 and P8 (10, 12).Cytoplasmic inclusion bodies, known as viroplasms or viral factories, are assumed to be the sites of replication of viruses in the family Reoviridae. After infecting insect vector cell monolayers (VCMs) in culture with RDV, Wei et al. (19) examined the generation of RDV particles in and at the periphery of such viroplasms. VCMs are also useful for studies of RGDV, allowing detailed analysis of the synchronous replication and multiplication of this virus (14). In order to identify the viroplasms in RGDV-infected VCMs, we examined the subcellular localization of Pns7, Pns12, P5, and RGDV particles by confocal immunofluorescence microscopy. Pns7 and Pns12 of RGDV correspond to Pns6 and Pns11, respectively, which are components of the viroplasm of RDV (12, 19). RGDV P5 is a counterpart of RDV P5, a core protein that locates inside the viroplasm in RDV-infected cells. We inoculated VCMs with RGDV, purified by the method reported in reference 15, at a multiplicity of infection (MOI) of 1; fixed them 48 h postinfection (p.i.); probed the cells with Pns7-, Pns12-, P5-, and viral-antigen-specific antibodies (11, 12) that had been conjugated to fluorescein isothiocyanate (FITC) (Sigma, St. Louis, MO) or rhodamine (Sigma); and examined them by confocal microscopy, as described previously (19). In RGDV-infected cells, Pns7, Pns12, and P5 were detected as punctate inclusions (Fig. (Fig.1).1). Immunostained viral antigens formed ringlike structures around the punctate inclusions. When the images were merged, Pns7, Pns12, and P5 were colocalized in the punctate inclusions, indicating that these proteins were constituents of the viral inclusions (Fig. (Fig.1).1). Our observations revealed the similar respective localizations of the corresponding nonstructural proteins, core proteins, and viral particles of two phytoreoviruses, RGDV and RDV, in infected cells. Thus, Pns7 and Pns12 of RGDV had attributes common to their functional counterparts—Pns6 and Pns11, respectively—of RDV (19). The core protein P5 was located inside the viroplasms, and the viral antigens were distributed at the periphery of the viroplasms. The results, together, suggest that RGDV and RDV exploit similar replication strategies. Specific fluorescence was not detected in noninfected cells after incubation with Pns7-, Pns12-, P5-, and viral-antigen-specific antibodies (data not shown).Open in a separate windowFIG. 1.Subcellular localization of Pns7, Pns12, and P5 of RGDV and viral antigens in RGDV-infected VCMs 48 h p.i. Arrowheads show ringlike profiles of viral antigens that surround viral inclusions, which have been immunostained with the Pns12-specific antibodies. Arrows show the fibrillar profiles of immunostained viral antigens. Bars, 5 μm.In addition to the viral location at the periphery of the viral inclusions visualized as immunostained Pns12 (Fig. (Fig.1),1), the antigens were distributed as bundles of fibrillar structures, a form not observed in RDV-infected cells. To analyze the entity of the bundles of fibrillar structures, VCMs on coverslips were inoculated with RGDV at an MOI of 1, fixed at 48 h p.i., and examined by electron microscopy (EM), as described previously (19). We observed viral particles of approximately 70 nm in diameter in close association with the free ends, as well as along the edges, of tubules of approximately 25 nm in diameter (Fig. 2A to D). The abundant bundles of tubules with closely associated viral particles were clearly in contact with the periphery of granular, electron-dense inclusions of 800 to 1,200 nm in diameter (Fig. (Fig.2B),2B), namely, viroplasms. The dimensions and appearance of the tubular structures resembled those of microtubules (Fig. (Fig.2C)2C) (17). Transverse sections of tubules revealed arrays of closed circles of approximately 25 nm in diameter, with viral particles attached directly or via a filament to the circumference (Fig. (Fig.2D2D).Open in a separate windowFIG. 2.Association of RGDV particles with microtubules. (A) Electron micrograph showing RGDV particles associated with microtubules in virus-infected VCMs 48 h p.i. Bar, 300 nm. VP, electron-dense inclusion. (B) Virus-associated microtubules in contact with the periphery of the electron-dense inclusion indicated by a white rectangle in panel A. Bar, 300 nm. (C) Viral particles along the edges of tubules of approximately 25 nm in diameter. Bar, 150 nm. (D) Transverse sections of arrays of closed circles of approximately 25 nm in diameter with viral particles attached to their circumference directly (arrow) or via a filament (arrowhead). Bar, 150 nm. (E) Confocal micrograph showing the association of viral particles with microtubules in virus-infected VCMs 48 h p.i. Microtubules were stained with α-tubulin-specific antibodies conjugated to FITC; viral particles were stained with viral-antigen-specific antibodies conjugated to rhodamine. Arrowheads indicate the ringlike organization of viral antigens. Arrows show the colocalization of fibrillar profiles of viral antigens with microtubules. The insets show ringlike and fibrillar profiles of immunostained viral antigens. The circular areas inside the ringlike structures are viroplasms. Bar, 5 μm.Our observations suggested that RGDV particles might attach to microtubules in infected cells. To examine this possibility, we inoculated VCMs with RGDV at an MOI of 1, fixed the cells 48 h p.i., immunostained them with α-tubulin-specific antibodies conjugated to FITC and with viral-antigen-specific antibodies conjugated to rhodamine, and analyzed them by confocal microscopy, as described previously (19). Viral antigens were visualized as ringlike and fibrillar structures (Fig. (Fig.2E).2E). Double immunostaining of the infected cells revealed that a network of microtubule-based filaments colocalized with most of the fibrillar structures that represented viral antigens, confirming the association of viral particles with the microtubule-like inclusions visualized by EM (Fig. (Fig.2A).2A). Nonspecific reactions were not detected with either of the stainings (data not shown). Our results suggested that RGDV particles, which assembled at the periphery of viroplasms, might be transported along microtubules. Due to the lack of RGDV infectious clones fused with green fluorescent protein and the effective gene transfection system for VCMs, we could not observe the trafficking of RGDV particles along microtubules in living cells.We then used three-dimensional (3-D) electron tomographic microscopy (ET) to reveal a new level of morphological detail about the association of RGDV with microtubules. To produce 3-D reconstructions of RGDV-infected cells, we fixed, embedded, and sectioned infected leafhopper cells as described previously (5). We chose a representative region that showed numerous RGDV particles close to bundles of microtubules for this novel tomographic analysis. A single-axis tilt series was collected manually from −60° to 60° with 2° increments using an H9500SD EM (Hitachi, Tokyo) operated at 200 kV. These tomographic data were recorded at a defocus of 3.6 μm on the TVIPS 2k × 2k charge-coupled-device camera (TVIPS, Gauting, Germany). Microscopic magnification of ×15,000, providing 1.28 nm/pixel, was enough to view the microtubules and virus particles following tomographic reconstruction of the tilt series using IMOD (7). As shown in the 3-D tomogram in Fig. Fig.3,3, most of the RGDV particles were bound to the edges of bundles of microtubules. The RGDV particles along the edges of microtubules were arrayed in an orderly but uncrowded manner (Fig. (Fig.3).3). Our ET analysis also revealed that some viral particles were linked to filaments of approximately 10 nm in diameter (Fig. (Fig.3B).3B). Morphologically, these filaments resembled vimentin intermediate filaments (4). In many lines of cultured cells, vimentin intermediate filaments partially overlap the microtubules, and there is evidence that the two filament systems interact (3, 9, 20). Unfortunately, vimentin-specific monoclonal antibodies from mouse and rabbit did not react specifically with our leafhopper cells (data not shown), but the nature of the intermediate filaments was apparent from their dimensions, intracellular location, and organization. Thus, our ET analysis indicated that RGDV particles were able to associate directly and/or via intermediate filaments with microtubules.Open in a separate windowFIG. 3.ET analysis showing the association of RGDV particles with microtubules either directly or via intermediate filaments. (A) Translucent representation of the reconstructed viruses lining up with microtubules. (B) Slice of the reconstructed volumes from the inset of A to show the association of RGDV particles with intermediate filaments (arrows). Bars, 150 nm.To examine the role of the microtubules for RGDV activity, we added a microtubule-disrupting agent, either nocodazole (Sigma) or colchicine (Sigma), 2 hours after inoculation of VCMs with RGDV at an MOI of 1 and then continued the incubation for a further 46 h. Cells were fixed 48 h p.i. and stained with α-tubulin-specific antibodies conjugated to FITC (Sigma) and viral-particle-specific antibodies conjugated to rhodamine, with subsequent confocal fluorescence microscopy, as described previously (19). We tested a range of drug concentrations in preliminary experiments (data not shown) and determined optimal concentrations. Treatment of infected cells with 10 μM nocodazole or 5 μg/ml colchicine resulted in the complete disassembly of microtubules, with the accumulation of ringlike structures exclusively and no fibrillar structures representative of viral antigens in the cytoplasm (Fig. (Fig.4A).4A). These ringlike aggregates of viral antigens were confirmed to surround viroplasms when the latter were immunostained for Pns12, as described above and shown in Fig. Fig.1.1. Nonspecific reactions were not detected with either staining (data not shown). These results suggest that RGDV particles multiply around the viroplasm but are unable to distribute along the microtubules in the presence of the chemicals.Open in a separate windowFIG. 4.(A) Effects of microtubule-disrupting agents on the formation of microtubules and fibrillar profiles of immunostained viral antigens. Bars, 5 μm. The insets show the ringlike profiles of immunostained viral antigens after treatment with inhibitors, suggesting that viral replication occurs in the presence of each inhibitor. (B) Effects of drugs on the production of cell-associated (gray bars) and extracellular (black bars) viruses in VCMs infected with RGDV. The error bars indicate standard deviations.During the process of infection, microtubules play important roles in viral entry, intracellular trafficking, and extracellular release (2, 8, 16). We next investigated the effects of the microtubule-disrupting agents on the production in and release of viruses from virus-infected cells by the method described previously (18). Nocodazole or colchicine was added 2 h after inoculation of VCMs with RGDV at an MOI of 1, and incubation was continued for a further 46 h. The extracellular medium and the cells were collected separately. The medium was centrifuged for 30 min at 15,000 × g, and the supernatant was collected. The cells were subjected to three cycles of freezing and thawing to release viral particles. The viral titer of each sample was determined, in duplicate, by the fluorescent focus assay as described previously (6), with VCMs and a magnification of ×10. As shown in Fig. Fig.4B,4B, nocodazole (20 μM) and colchicine (10 μg/ml) caused a fivefold reduction in the number of released viruses, compared to that from untreated control infected cells. In contrast, each inhibitor at the selected dose failed to significantly reduce the titer of cell-associated viruses (less then 5% compared to that from untreated control). These results suggest that the inhibitors impeded the release of viruses into the medium without affecting viral production in infected cells. We do not yet understand why the viral titer was not elevated in drug-treated cells from which viral release was inhibited. However, our data show clearly that disruption of microtubules directly inhibited the release of mature viral particles from infected cells.In conclusion, EM, ET, immunofluorescence staining, and experiments with two inhibitors support the hypothesis that the transport of RGDV from viroplasms to the plasma membrane and into the medium is dependent on microtubules. In the case of RDV, vesicular compartment-containing viral particles that locate adjacent to the viroplasms were considered to play an important role in the transport and release of the virus from the viroplasm to the culture medium in infected VCMs (18). On the other hand, a fibrillar structure (Fig. (Fig.11 and and2),2), not observed in RDV-infected cells, was considered to function in the trafficking of RGDV from viroplasm into the culture medium (Fig. (Fig.4)4) in the present study. RGDV and RDV, both members of the Phytoreovirus genus, have some common biological and biochemical properties but are distinct from each other (13). For example, viruses are restricted to phloem-related cells in RGDV-infected plants but distributed in many types of cells in RDV-infected plants, and a P2 protein with a function to adsorb to and/or penetrate into insect vector cells is present in RGDV and absent in RDV in particles purified using carbon tetrachloride. The present molecular cytopathological study revealed one more difference between the viruses: they have different means for transporting and releasing infectious particles to the cell exterior. The presence of such a molecular mechanism may accelerate the secondary infections by the viruses in infected vector insects, and the high propagation speed would allow the viruses to complete infection cycles through insects and plants.  相似文献   

14.
15.
16.
The Merkel cell polyomavirus (MCPyV) was identified recently in human Merkel cell carcinomas, an aggressive neuroendocrine skin cancer. Here, we identify a putative host cell receptor for MCPyV. We found that recombinant MCPyV VP1 pentameric capsomeres both hemagglutinated sheep red blood cells and interacted with ganglioside GT1b in a sucrose gradient flotation assay. Structural differences between the analyzed gangliosides suggest that MCPyV VP1 likely interacts with sialic acids on both branches of the GT1b carbohydrate chain. Identification of a potential host cell receptor for MCPyV will aid in the elucidation of its entry mechanism and pathophysiology.Members of the polyomavirus (PyV) family, including simian virus 40 (SV40), murine PyV (mPyV), and BK virus (BKV), bind cell surface gangliosides to initiate infection (2, 8, 11, 15). PyV capsids are assembled from 72 pentamers (capsomeres) of the major coat protein VP1, with the internal proteins VP2 and VP3 buried within the capsids (7, 12). The VP1 pentamer makes direct contact with the carbohydrate portion of the ganglioside (10, 12, 13) and dictates the specificity of virus interaction with the cell. Gangliosides are glycolipids that contain a ceramide domain inserted into the plasma membrane and a carbohydrate domain that directly binds the virus. Specifically, SV40 binds to ganglioside GM1 (2, 10, 15), mPyV binds to gangliosides GD1a and GT1b (11, 15), and BKV binds to gangliosides GD1b and GT1b (8).Recently, a new human PyV designated Merkel cell PyV (MCPyV) was identified in Merkel cell carcinomas, a rare but aggressive skin cancer of neuroendocrine origin (3). It is as yet unclear whether MCPyV is the causative agent of Merkel cell carcinomas (17). A key to understanding the infectious and transforming properties of MCPyV is the elucidation of its cellular entry pathway. In this study, we identify a putative host cell receptor for MCPyV.Because an intact infectious MCPyV has not yet been isolated, we generated recombinant MCPyV VP1 pentamers in order to characterize cellular factors that bind to MCPyV. VP1 capsomeres have been previously shown to be equivalent to virus with respect to hemagglutination properties (4, 16), and the atomic structure of VP1 bound to sialyllactose has demonstrated that the capsomere is sufficient for this interaction (12, 13). The MCPyV VP1 protein (strain w162) was expressed and purified as described previously (1, 6). Briefly, a glutathione S-transferase-MCPyV VP1 fusion protein was expressed in Escherichia coli and purified using glutathione-Sepharose affinity chromatography. The fusion protein was eluted using glutathione and cleaved in solution with thrombin. The thrombin-cleaved sample was then rechromatographed on a second glutathione-Sepharose column to remove glutathione transferase and any uncleaved protein. The unbound VP1 was then chromatographed on a P-11 phosphocellulose column, and peak fractions eluting between 400 and 450 mM NaCl were collected. The purified protein was analyzed by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE), followed by Coomassie blue staining (Fig. (Fig.1A,1A, left) and immunoblotting using an antibody (I58) that generally recognizes PyV VP1 proteins (Fig. (Fig.1A,1A, right) (9). Transmission electron microscopy (Philips CM10) analysis confirmed that the purified recombinant MCPyV VP1 formed pentamers (capsomeres), which did not assemble further into virus-like particles (Fig. (Fig.1B).1B). In an initial screening of its cell binding properties, we tested whether the MCPyV VP1 pentamers hemagglutinated red blood cells (RBCs). The MCPyV VP1 pentamers were incubated with sheep RBCs and assayed as previously described (5). SV40 and mPyV recombinant VP1 pentamers served as negative and positive controls, respectively. We found that MCPyV VP1 hemagglutinated the RBCs with the same efficiency as mPyV VP1 (protein concentration/hemagglutination unit) (Fig. (Fig.1C,1C, compare rows B and C from wells 1 to 11), suggesting that MCPyV VP1 engages a plasma membrane receptor on the RBCs. The recombinant murine VP1 protein used for comparison was from the RA strain, a small plaque virus (4). Thus, MCPyV VP1 has the hemagglutination characteristics of a small plaque mPyV (12, 13).Open in a separate windowFIG. 1.Characterization of MCPyV VP1. Recombinant MCPyV VP1 forms pentamers and hemagglutinates sheep RBCs. (A) Coomassie blue-stained SDS-PAGE and an immunoblot of the purified recombinant MCPyV VP1 protein are shown. Molecular mass markers are indicated. (B) Electron micrograph of the purified MCPyV VP1. MCPyV VP1 (shown in panel A) was diluted to 100 μg/ml and absorbed onto Formvar/carbon-coated copper grids. Samples were washed with phosphate-buffered saline, stained with 1% uranyl acetate, and visualized by transmission electron microscopy at 80 kV. Bar = 20 nm. (C) Sheep RBCs (0.5%) were incubated with decreasing concentrations of purified recombinant SV40 VP1 (row A), mPyV VP1 (row B), and MCPyV VP1 (row C). Wells 1 to 11 contain twofold serial dilutions of protein, starting at 2 μg/ml (well 1). Well 12 contains buffer only and serves as a negative control. Well 7 (rows B and C) corresponds to 128 hemagglutination units per 2 μg/ml VP1 protein.To characterize the chemical nature of the putative receptor for MCPyV, total membranes from RBCs were purified as described previously (15). The plasma membranes (30 μg) were incubated with MCPyV VP1 (0.5 μg) and floated on a discontinuous sucrose gradient (15). After fractionation, the samples were analyzed by SDS-PAGE, followed by immunoblotting with I58. VP1 was found in the bottom of the gradient in the absence of the plasma membranes (Fig. (Fig.2A,2A, first panel). In the presence of plasma membranes, a fraction of the VP1 floated to the middle of the gradient (Fig. (Fig.2A,2A, second panel), supporting the hemagglutination results that suggested that MCPyV VP1 binds to a receptor on the plasma membrane.Open in a separate windowFIG. 2.MCPyV VP1 binds to a protease-resistant, sialic acid-containing receptor on the plasma membrane. (A) Purified recombinant MCPyV VP1 was incubated with or without the indicated plasma membranes. The samples were floated in a discontinuous sucrose gradient, and the fractions were collected from the top of the gradient, subjected to SDS-PAGE, and immunoblotted with the anti-VP1 antibody I58. (B) Control and proteinase K-treated plasma membranes were subjected to SDS-PAGE, followed by Coomassie blue staining. (C) HeLa cells treated with proteinase K (4 μg/ml) were incubated with MCPyV at 4°C, and the resulting cell lysate was probed for the presence of MCPyV VP1. (D) As described in the legend to panel C, except 293T cells were used. (E) Purified MCPyV VP1 was incubated with plasma membranes pretreated with or without α2-3,6,8 neuraminidase and analyzed as described in the legend to panel A.To determine whether the receptor is a protein or a lipid, plasma membrane preparations (30 μg) were incubated with proteinase K (Sigma), followed by analysis with SDS-PAGE and Coomassie blue staining. Under these conditions, the majority of the proteins in the plasma membranes were degraded by the protease (Fig. (Fig.2B,2B, compare lanes 1 and 2). Despite the lack of proteins, the proteinase K-treated plasma membranes bound MCPyV VP1 as efficiently as control plasma membranes (Fig. (Fig.2A,2A, compare the second and third panels), demonstrating that MCPyV VP1 interacts with a protease-resistant receptor. The absence of VP1 in the bottom fraction in Fig. Fig.2A2A (third panel) is consistent with the fact that the buoyant density of the membranes is lowered by proteolysis. Of note, a similar result was seen with binding of the mPyV to the plasma membrane (15). Binding of MCPyV to the cell surface of two human tissue culture cells (i.e., HeLa and 293T) was also largely unaffected by pretreatment of the cells with proteinase K (Fig. 2C and D, compare lanes 1 and 2), further indicating that a nonproteinaceous molecule on the plasma membrane engages the virus.We next determined whether the protease-resistant receptor contains a sialic acid modification. Plasma membranes (10 μg) were incubated with a neuraminidase (α2-3,6,8 neuraminidase; Calbiochem) to remove the sialic acid groups. In contrast to the control plasma membranes, the neuraminidase-treated membranes did not bind MCPyV VP1 (Fig. (Fig.2E,2E, compare first and second panels), indicating that the MCPyV receptor includes a sialic acid modification.Gangliosides are lipids that contain sialic acid modifications. We asked if MCPyV VP1 binds to gangliosides similar to other PyV family members. The structures of the gangliosides used in this analysis (gangliosides GM1, GD1a, GD1b, and GT1b) are depicted in Fig. Fig.3A.3A. To assess a possible ganglioside-VP1 interaction, we employed a liposome flotation assay established previously (15). When liposomes (consisting of phosphatidyl-choline [19 μl of 10 mg/ml], -ethanolamine [5 μl of 10 mg/ml], -serine [1 μl of 10 mg/ml], and -inositol [3 μl of 10 mg/ml]) were incubated with MCPyV VP1 and subjected to the sucrose flotation assay, the VP1 remained in the bottom fraction (Fig. (Fig.3B,3B, first panel), indicating that VP1 does not interact with these phospholipids. However, when liposomes containing GT1b (1 μl of 1 mM), but not GM1 (1 μl of 1 mM) or GD1a (1 μl of 1 mM), were incubated with MCPyV VP1, the vesicles bound this VP1 (Fig. (Fig.3B).3B). A low level of virus floated partially when incubated with liposomes containing GD1b (Fig. (Fig.3B),3B), perhaps reflecting a weak affinity between MCPyV and GD1b. Importantly, MCPyV binds less efficiently to neuraminidase-treated GT1b-containing liposomes than to GT1b-containing liposomes (Fig. (Fig.3B,3B, sixth panel), suggesting that the GT1b sialic acids are involved in virus binding. This finding is consistent with the ability of neuraminidase to block MCPyV binding to the plasma membrane (Fig. (Fig.2E).2E). The level of virus flotation observed in the neuraminidase-treated GT1b-containing liposomes is likely due to the inefficiency of the neuraminidase reaction with a high concentration of GT1b used to prepare the vesicles.Open in a separate windowFIG. 3.MCPyV VP1 binds to ganglioside GT1b. (A) Structures of gangliosides GM1, GD1a, GD1b, and GT1b. The nature of the glycosidic linkages is indicated. (B) Purified MCPyV VP1 protein was incubated with liposomes only or with liposomes containing the indicated gangliosides. The samples were analyzed as described in the legend to Fig. Fig.2A.2A. Where indicated, GT1b-containing liposomes were pretreated with α2-3,6,8 neuraminidase and analyzed subsequently for virus binding. (C to E) The indicated viruses were incubated with liposomes and analyzed as described in the legend to panel B.As controls, GM1-containing liposomes bound SV40 (Fig. (Fig.3C),3C), GD1a-containing liposomes bound mPyV (Fig. (Fig.3D),3D), and GD1b-containing liposomes bound BKV (Fig. (Fig.3E),3E), demonstrating that the liposomes were functionally intact. We note that, while all of the MCPyV VP1 floated when incubated with liposomes containing GT1b (Fig. (Fig.3B,3B, sixth panel), a significant fraction of SV40, mPyV, and BKV VP1 remained in the bottom fraction despite being incubated with liposomes containing their respective ganglioside receptors (Fig. 3C to E, second panels). This result is likely due to the fact that in contrast to MCPyV, which are assembled as pentamers (Fig. (Fig.1B),1B), the SV40, mPyV, and BKV used in these experiments are fully assembled particles: their larger and denser nature prevents efficient flotation. Nonetheless, we conclude that MCPyV VP1 binds to ganglioside GT1b efficiently.The observation that GD1a does not bind to MCPyV VP1 suggests that the monosialic acid modification on the right branch of GT1b (Fig. (Fig.3A)3A) is insufficient for binding. Similarly, the failure of GD1b to bind MCPyV VP1 suggests that the sialic acid on the left arm of GT1b is necessary for binding. Together, these observations suggest that MCPyV VP1 interacts with sialic acids on both branches of GT1b (Fig. (Fig.4).4). A recent structure of SV40 VP1 in complex with the sugar portion of GM1 (10) demonstrated that although SV40 VP1 binds both branches of GM1 (Fig. (Fig.4),4), only a single sialic acid in GM1 is involved in this interaction. In the case of mPyV, structures of mPyV VP1 in complex with different carbohydrates (12, 13) revealed that the sialic acid-galactose moiety on the left branch of GD1a (and GT1b) is sufficient for mPyV VP1 binding (Fig. (Fig.4).4). Although no structure of BKV in complex with the sugar portion of GD1b (or GT1b) is available, in vitro binding studies (8) have suggested that the disialic acid modification on the right branch of GD1b (and GT1b) is responsible for binding BKV VP1 (Fig. (Fig.4).4). Thus, it appears that the unique feature of the MCPyV VP1-GT1b interaction is that the sialic acids on both branches of this ganglioside are likely involved in capsid binding.Open in a separate windowFIG. 4.A potential model of the different VP1-ganglioside interactions (see the text for discussion).The identification of a potential cellular receptor for MCPyV will facilitate the study of its entry mechanism. An important issue for further study is to determine whether MCPyV targets Merkel cells preferentially, and if so, whether GT1b is found in higher levels in these cells to increase susceptibility.  相似文献   

17.
The sheath of the Serratia entomophila antifeeding prophage, which is pathogenic to the New Zealand grass grub Costelytra zealandica, is a 3-fold helix formed by a 4-fold symmetric repeating motif disposed around a helical inner tube. This structure, determined by electron microscopy and image processing, is distinct from that of the other known morphologically similar bacteriophage sheaths.The antifeeding prophage (Afp) of Serratia entomophila and Serratia proteamaculans is a naturally occurring virus tail-like structure which delivers a putative toxin molecule that leads to starvation of the New Zealand grass grub Costelytra zealandica (5). Afp is composed of 18 different gene products (molecular masses of 6.5 to 263 kDa). The first 16 open reading frames have orthologues (Photorhabdus virulence cassettes [PVC]) in the insecticidal bacterium Photorhabdus luminescens TTO1 genome (5). Afp and PVCs morphologically resemble a typical R-type bacteriocin (6, 12, 16) However, Afp is the only known phage tail-like protein complex that is not a bacteriocin-protein complex of considerable medical relevance that targets the same or closely related bacterial strains (1, 3, 8, 12). The major component of Afp is a contractile cylindrical outer sheath encasing an inner tube speculated to house the toxin molecule (6). A dome-shaped “head” defines one extremity of the tube, while the other end is attached to a “bell-shaped” structure with a base morphologically similar to the base plate of the T4 bacteriophage tail (9).Transmission electron micrographs of two-dimensional (2D) projections of negatively stained (Fig. (Fig.11 A) or frozen-hydrated and vitrified (Fig. (Fig.1B)1B) recombinant Afp particles (see Fig. S1 in the supplemental material) were used for computational image analysis. A globally averaged image of the Afp particle in the major configuration (called E here) (Fig. (Fig.1C),1C), generated using negatively stained specimens, clearly distinguished the morphologies of the various constituent structural parts. Thus, the cylindrical sheath appears to be formed by a periodic structure harboring a distinctive, inverted-V-shaped feature. A minor population of Afp particles displays an alternate configuration (called C here) where, concomitant with contraction of the sheath (averaged axial compression of ∼52% [see Table S1 in the supplemental material]), the inner tube, shorn off the bell-shaped structure, is revealed (Fig. (Fig.1A)1A) (6). Several other bacteriocins undergo such a high degree of compression, which has been characterized in detail for the tail sheath of bacteriophage T4 (9). We also generated individual global averages for the periodic sheath structure, for the bell-shaped structure, and for the inner tube (Fig. (Fig.1C)1C) which provide more accurate dimensions of these different sections (see Table S1 in the supplemental material) than those reported earlier (6).Open in a separate windowFIG. 1.(A) Electron micrograph of a negatively stained preparation of partially purified recombinant Afp particles. The gray, white, and black arrows point to an Afp particle in the extended (E) configuration, to an Afp particle with the sheath contracted, exposing the inner tube in the contracted (C) configuration, and to an inner tube with the bell-shaped structure attached at one end, respectively. (B) Cryoelectron micrograph recorded from a preparation similar to that shown in panel A. (C) Globally averaged images of Afp particles (3,026 images) in the E state and the three distinguishable parts visualized by negative staining. Bars, 200 Å. (D) Averaged power spectrum of the sheath of vitrified Afp particles. The black and white arrows indicate reflections delineating the axial rise (1/78.5 Å) and the helical pitch (1/118 Å), respectively. In panels A and C, lighter regions represent protein and the contrast is reversed in panel B. Global averages were created using classalign2 of the EMAN suite (10) and visualized in bshow (4).For a better insight, we determined the 3D structure of the central periodic section of the Afp particle in the E state. A global power spectrum derived from the cryoimages established the structure to be helical with a clear first meridional reflection at 1/78.5 Å and the first, strongest nonmeridional reflection at 1/118 Å (Fig. (Fig.1D).1D). These reflections correspond to the axial rise (Δx) and the pitch of the helix, respectively, and reflect a turn angle (Δψ) of about ±240° (32 helix) for the repeating motif. The correct sign, i.e., the hand, of the helix remains to be determined. Computationally excised overlapping segments of this helical section from images of vitrified and negatively stained Afp particles were subjected to 3D reconstruction using the iterative helical real-space reconstruction (IHRSR) algorithm (2) using the determined helical parameters (see Fig. S2 in the supplemental material). After a few iterations, the presence of an in-plane 4-fold symmetry (C4) was apparent in the density map (see Fig. S2 in the supplemental material), which was then imposed in the subsequent reconstruction exercises. However, no stable solution was forthcoming, even after many (e.g., 30) iterative cycles. This is generally indicative of the presence of heterogeneity in the form of variations in helix translation and/or twist angle (15) in the structure. As a first step, we focused our attention on the pitch value, and following classification (see Fig. S3 in the supplemental material), we found that the majority of the image segments correspond to a pitch of 120 Å. These segments were then selected out of the full data set and led to a stable and refined 3D reconstruction. We also obtained very comparable results for the helical section when images of negatively stained Afp sheath sections were used, thus supporting our computational approach (see Fig. S4 in the supplemental material) and general conclusions about the E state described below.Figure Figure22 A is an isosurface representation of the density map of the helical Afp sheath in the E state calculated at ∼21.5-Å resolution (see Fig. S5 in the supplemental material). To the best of our knowledge, a 4-fold rotational symmetry has not been seen for any other contractile T4 bacteriophage taillike structure, which points to the unique architecture of the Afp sheath. A power spectrum generated using the 2D projection from the final density map, compared to the experimental global power spectrum (Fig. (Fig.2D),2D), showed strong agreement, confirming the fidelity of the computational image analysis. The density map displays protein layers, ∼80 Å thick, that are stacked on each other in a periodic fashion. The uneven outer surface of the sheath is perforated and decorated with ∼35-Å protrusions. When rendered with a raised threshold, a characteristic feature of the map is a contiguous, high-electron-density region having an inverted-Y-shaped structure (Fig. 2C and E; see Fig. S4 in the supplemental material). At the modest ∼21.5-Å resolution, the boundary of the repeating subunit cannot be defined. A 25 ± 3-Å-wide central lumen is seen clearly when viewed along the helix axis (Fig. (Fig.2B)2B) and likely represents the pore of the inner tube (see also below). Using scanning transmission electron microscopy (STEM) (see Fig. S6 in the supplemental material), we estimated the averaged molecular mass of the central helical section of an Afp particle to be 9.8 ± 0.4 kDa/Å (Fig. (Fig.3)3) (14) and that of only the inner tube to be ∼2.5 kDa/Å, based on a relatively small pool of such images. These values translate to a mass contribution of approximately 145 kDa of the subunit whose periodic arrangement forms the outer component of the sheath (i.e., excluding the inner tube) (see Fig. S6 in the supplemental material). This value is not very different from the cumulative mass of the different proteins, i.e., homologous afp2, afp3, and afp4, thought to be involved in Afp sheath formation (5) (see Fig. S7 in the supplemental material).Open in a separate windowFIG. 2.Orthogonal isosurface rendering, at 1 σ (standard deviation) of the computed ∼21.5-Å density map of the helical sheath of the vitrified Afp particle viewed normal (A) and parallel (B) to the helix. The images were generated using the software package CHIMERA (13). The arrow indicates a surface perforation. (C) The Afp density map rendered at 3.5 σ to highlight the largest contiguous high-electron-density regions; one circumscribed by a black ellipse is computationally extracted and shown in panel E. (D) Comparison of the experimental, averaged power spectrum of the helical sheath of Afp (left) with that computed (right) from the 2D projection of the calculated density map. (F) Global average of the inner tube of the Afp particle and a plot of the surface density variation (scaled from 0 to 1) along the helical (y) axis. The dimension along the tube is plotted on the x axis.Open in a separate windowFIG. 3.(A) Dark-field micrograph of a freeze-dried, unstained preparation of Afp particles used in STEM measurements. An Afp particle in the E state, an Afp inner tube with the attached bell-shaped structure, and a tobacco mosaic virus particle, used as a calibration standard, are marked by the arrowhead and the gray and white arrows, respectively. (B) Histogram plot of the measured distributions of mass per unit length corresponding to the uniform periodic section of the Afp particles overlaid with a fitted Gaussian curve produced by using the ORIGIN6 software package.A paucity of images of the C state (∼5% of the complete data set) precluded a full, refined 3D reconstruction, but based on the available 2,774 overlapping image segments of the isolated inner tube, a global average was calculated. A plot of contrast variation (Fig. (Fig.2F)2F) indicates that the surface is characterized by ∼40-Å spaced elevated crests and invaginated grooves, in agreement with the calculated axial rise of ∼39 Å for the subunit (see Fig. S8 in the supplemental material) comprising the tube. Based on these preliminary results, it appears that the helical symmetry of the inner tube is markedly different from that of the sheath.Our observation that the pitch of the helix in the E state can vary by as much as ∼50 Å attests to the flexible nature of the sheath, which is required for compressibility and may be facilitated by the somewhat porous nature of the sheath (Fig. (Fig.2).2). Preliminary deductions (data not shown) based on a small pool of images of the C state suggest profound rearrangement of the elements of the sheath. How that translates to extrusion of the toxin remains to be revealed.   相似文献   

18.
The bicistronic groESL operon, encoding the Hsp60 and Hsp10 chaperonins, was cloned into an integrative expression vector, pFPN, and incorporated at an innocuous site in the Anabaena sp. strain PCC7120 genome. In the recombinant Anabaena strain, the additional groESL operon was expressed from a strong cyanobacterial PpsbA1 promoter without hampering the stress-responsive expression of the native groESL operon. The net expression of the two groESL operons promoted better growth, supported the vital activities of nitrogen fixation and photosynthesis at ambient conditions, and enhanced the tolerance of the recombinant Anabaena strain to heat and salinity stresses.Nitrogen-fixing cyanobacteria, especially strains of Nostoc and Anabaena, are native to tropical agroclimatic conditions, such as those of Indian paddy fields, and contribute to the carbon (C) and nitrogen (N) economy of these soils (22, 30). However, their biofertilizer potential decreases during exposure to high temperature, salinity, and other such stressful environments (1). A common target for these stresses is cellular proteins, which are denatured and inactivated during stress, resulting in metabolic arrest, cessation of growth, and eventually loss of viability. Molecular chaperones play a major role in the conformational homeostasis of cellular proteins (13, 16, 24, 26) by (i) proper folding of nascent polypeptide chains; (ii) facilitating protein translocation and maturation to functional conformation, including multiprotein complex assembly; (iii) refolding of misfolded proteins; (iv) sequestering damaged proteins to aggregates; and (v) solubilizing protein aggregates for refolding or degradation. Present at basal levels under optimum growth conditions in bacteria, the expression of chaperonins is significantly enhanced during heat shock and other stresses (2, 25, 32).The most common and abundant cyanobacterial chaperones are Hsp60 proteins, and nitrogen-fixing cyanobacteria possess two or more copies of the hsp60 or groEL gene (http://genome.kazusa.or.jp/cyanobase). One occurs as a solitary gene, cpn60 (17, 21), while the other is juxtaposed to its cochaperonin encoding genes groES and constitutes a bicistronic operon groESL (7, 19, 31). The two hsp60 genes encode a 59-kDa GroEL and a 61-kDa Cpn60 protein in Anabaena (2, 20). Both the Hsp60 chaperonins are strongly expressed during heat stress, resulting in the superior thermotolerance of Anabaena, compared to the transient expression of the Hsp60 chaperonins in Escherichia coli (20). GroEL and Cpn60 stably associate with thylakoid membranes in Anabaena strain PCC7120 (14) and in Synechocystis sp. strain PCC6803 (15). In Synechocystis sp. strain PCC6803, photosynthetic inhibitors downregulate, while light and redox perturbation induce cpn60 expression (10, 25, 31), and a cpn60 mutant exhibits a light-sensitive phenotype (http://genome.kazusa.or.jp/cyanobase), indicating a possible role for Cpn60 in photosynthesis. GroEL, a lipochaperonin (12, 28), requires a cochaperonin, GroES, for its folding activity and has wider substrate selectivity. In heterotrophic nitrogen-fixing bacteria, such as Klebsiella pneumoniae and Bradyrhizobium japonicum, the GroEL protein has been implicated in nif gene expression and the assembly, stability, and activity of the nitrogenase proteins (8, 9, 11).Earlier work from our laboratory demonstrated that the Hsp60 family chaperonins are commonly induced general-stress proteins in response to heat, salinity, and osmotic stresses in Anabaena strains (2, 4). Our recent work elucidated a major role of the cpn60 gene in the protection from photosynthesis and the nitrate reductase activity of N-supplemented Anabaena cultures (21). In this study, we integrated and constitutively overexpressed an extra copy of the groESL operon in Anabaena to evaluate the importance and contribution of GroEL chaperonin to the physiology of Anabaena during optimal and stressful conditions.Anabaena sp. strain PCC7120 was photoautotrophically grown in combined nitrogen-free (BG11) or 17 mM NaNO3-supplemented (BG11+) BG11 medium (5) at pH 7.2 under continuous illumination (30 μE m−2 s−1) and aeration (2 liters min−1) at 25°C ± 2°C. Escherichia coli DH5α cultures were grown in Luria-Bertani medium at 37°C at 150 rpm. For E. coli DH5α, kanamycin and carbenicillin were used at final concentrations of 50 μg ml−1 and 100 μg ml−1, respectively. Recombinant Anabaena clones were selected on BG11+ agar plates supplemented with 25 μg ml−1 neomycin or in BG11 liquid medium containing 12.5 μg ml−1 neomycin. The growth of cyanobacterial cultures was estimated either by measuring the chlorophyll a content as described previously (18) or the turbidity (optical density at 750 nm). Photosynthesis was measured as light-dependent oxygen evolution at 25 ± 2°C by a Clark electrode (Oxy-lab 2/2; Hansatech Instruments, England) as described previously (21). Nitrogenase activity was estimated by acetylene reduction assays, as described previously (3). Protein denaturation and aggregation were measured in clarified cell extracts containing ∼500 μg cytosolic proteins treated with 100 μM 8-anilino-1-naphthalene sulfonate (ANS). The pellet (protein aggregate) was solubilized in 20 mM Tris-6 M urea-2% sodium dodecyl sulfate (SDS)-40 mM dithiothreitol for 10 min at 50°C. The noncovalently trapped ANS was estimated using a fluorescence spectrometer (model FP-6500; Jasco, Japan) at a λexcitation of 380 nm and a λemission of 485 nm, as described previously (29).The complete bicistronic groESL operon (2.040 kb) (GenBank accession no. FJ608815) was PCR amplified from PCC7120 genomic DNA using specific primers (Table (Table1)1) and the amplicon cloned into the NdeI-BamHI restriction sites of plasmid vector pFPN, which allows integration at a defined innocuous site in the PCC7120 genome and expression from a strong cyanobacterial PpsbA1 promoter (6). The resulting construct, designated pFPNgro (Table (Table1),1), was electroporated into PCC7120 using an exponential-decay wave form electroporator (200 J capacitive energy at a full charging voltage of 2 kV; Pune Polytronics, Pune, India), as described previously (6). The electroporation was carried out at 6 kV cm−1 for 5 ms, employing an external autoclavable electrode with a 2-mm gap. The electroporation buffer contained high concentrations of salt (10 mM HEPES, 100 mM LiCl, 50 mM CaCl2), as have been recommended for plant cells (23) and other cell types (27). The electrotransformants, selected on BG11+ agar plates supplemented with 25 μg ml−1 neomycin by repeated subculturing for at least 25 weeks to achieve complete segregation, were designated AnFPNgro.

TABLE 1.

Plasmids, strains, and primers used in this study
Plasmid, strain, or primerFeature or sequenceaSource or reference
Plasmids
    pFPNIntegrative expression vector6
    pFPNgropFPN with groESL operonThis study
Strains
    An7120Wild-type Anabaena sp. strain PCC7120R. Haselkorn
    AnFPNgroGroESL-overexpressing AnabaenaThis study
Primers
    groESLfwd5′-GGA ATT CCA TAT GGC AGC AGT ATC TCT AAG-3′This study
    groESLrev5′-CGC GGA TCC TTA GTA ATC GAA GTC ACC GCC-3′This study
    PpsbA1fwd5′-GAG CTG CAG GGA TTC CCA AAG ATA GGG-3′6
    PpsbA1rev5′-CTC GGA TCC CCA TAT GTT TTT ATG ATT GCT TTG-3′6
Open in a separate windowaThe underlined nucleotides in the primer sequences represent the incorporated restriction endonuclease sites.The transfer of pFPNgro to PCC7120 resulted in the integration of an extra copy of groESL (PpsbA1-groESL) into the PCC7120 genome. PCR amplification (Fig. (Fig.1I)1I) with the PpsbA1 forward and groESL reverse primer pairs showed the additional copy of groEL juxtaposed downstream to the PpsbA1 promoter (lane 6) in the recombinant Anabaena strain, while the native groESL operon found in the wild-type strain (lane 3) remained intact in the AnFPNgro strain (lane 5).Open in a separate windowFIG. 1.Integration and constitutive expression of an additional groESL operon in Anabaena strain PCC7120. (I) Integration of an additional groESL operon in the PCC7120 genome. The electrophoretogram shows the transfer and integration of PpsbA1-groESL in strain AnFPNgro. Lane 1, 1-kb DNA marker; lane 2, PCR control template without primer; lane 3, PCR product from wild-type Anabaena using the groESLfwd and groESLrev primers; lane 4, PCR product from PCC7120 using the PpsbA1fwd and groESLrev primers; lane 5, PCR product from AnFPNgro using the groESLfwd and groESLrev primers; lane 6, PCR product from AnFPNgro using the PpsbA1fwd and groESLrev primers. (II) Expression of the groESL operon in the wild-type and recombinant Anabaena strains during stress. PCC7120 (An7120) and AnFPNgro were grown for 3 days and then subjected to either heat stress (42°C) for 4 h (A and A′) or salinity stress (150 mM NaCl) for 3 days (B and B′). GroEL levels were estimated by Western blotting of 10% SDS-polyacrylamide gel electrophoresis-resolved whole-cell proteins, followed by immunodetection using anti-AnGroEL antiserum and densitometry (A and B). Panels A′ and B′ depict SDS-polyacrylamide gel electrophoresis-resolved and Coomassie blue-stained proteins to show equal sample loading. Various lanes contained protein samples under unstressed-control (U), heat (H), or salt (S) stress conditions. Numbers below panels A and B show GroEL quantitation by densitometry.Under normal growth conditions, the recombinant AnFPNgro cells expressed about 8.7- to 9.9-fold higher levels of GroEL protein than that detected in the PCC7120 cells (Fig. 1II), indicating a strong constitutive expression of the GroEL protein from the PpsbA1 promoter. In PCC7120, the wild-type copy of the GroEL protein was induced by both heat shock (Fig. 1IIA, lane 2) and salt stress (Fig. 1IIB, lane 2). GroEL levels in the recombinant strain were found to be about 2.5-fold higher under heat stress (Fig. 1IIA, lane 4) and approximately 1.7-fold higher under salinity stress (Fig. 1IIB, lane 4) than that expressed by PCC7120 under these stresses (Fig. 1IIA and IIB, lanes 2). The exposure of AnFPNgro cells to heat stress resulted in a further increase of approximately sixfold in GroEL levels (Fig. 1IIA, lane 4), while salt stress enhanced GroEL levels by approximately threefold (Fig. 1IIB, lane 4), compared to the constitutively expressed GroEL level in this strain (Fig. 1IIA and IIB, lanes 3). The constitutive expression of GroEL protein in AnFPNgro under ambient conditions (Fig. 1IIA and IIB, lanes 3) was from the PpsbA1 promoter (Fig. (Fig.1I,1I, lane 6). We assume that the additional increase in GroEL levels observed under heat and salt stress (Fig. 1IIA and IIB, lanes 4) was due to the native stress-induced groESL operon, functional from its own promoter.The diazotrophically grown PCC7120 did not grow during prolonged exposure to heat stress (42°C) (Fig. (Fig.2A)2A) and showed poor growth during salinity stress (150 mM) (Fig. (Fig.2B).2B). Salinity stress was particularly severe for photosynthetic pigments in PCC7120 and bleached the cells (data not shown). In contrast, the recombinant strain AnFPNgro showed a higher content of major photosynthetic pigments (Fig. (Fig.2C)2C) and presented a healthier blue-green phenotype (data not included). Strain AnFPNgro also showed better growth than wild-type PCC7120, both under unstressed and stressed conditions (Fig. 2A and B).Open in a separate windowFIG. 2.Effect of groESL overexpression on thermotolerance and salinity tolerance of diazotrophically grown Anabaena strains. (A) Growth (measured as chlorophyll a content) of strains during prolonged exposure to 42°C. (B) Growth (turbidity measured at an optical density at 750 nm) during prolonged exposure to 150 mM NaCl. (C) Absorption spectra of a dilute suspension of whole filaments after 7 days of exposure to various NaCl concentrations.The photosynthetic activity decreased with time during heat stress in PCC7120 but was maintained at comparatively higher levels in AnFPNgro cells (Fig. (Fig.3A)3A) than in PCC7120. The dinitrogenase activity in PCC7120 was severely inhibited after 4 h of heat stress (Fig. (Fig.3B).3B). In contrast, the dinitrogenase activity of the recombinant strain (AnFPNgro) was about 1.5-fold higher than PCC7120 under ambient conditions (25°C ± 2°C, no NaCl) and more than 3-fold higher than that of PCC7120 after 4 h of heat stress (Fig. (Fig.3B).3B). Prolonged exposure to salinity stress inhibited photosynthesis and nitrogen fixation in PCC7120 (Fig. 3C and D). However, strain AnFPNgro displayed significant protection of these activities, possibly due to overexpressed GroES/GroEL proteins. The recombinant strain (AnFPNgro) exhibited much-reduced protein aggregation after 4 h of heat stress or after prolonged exposure (10 days) to salinity stress than PCC7120 (Fig. (Fig.44).Open in a separate windowFIG. 3.Effect of groESL overexpression on photosynthesis and nitrogen fixation in Anabaena. Photosynthesis (A and C) and nitrogenase activity (B and D) in wild-type Anabaena strain PCC7120 (An7120) and recombinant AnFPNgro strains exposed to heat stress for 10 days (A) or 4 h (B) or to salinity stress (150 mM) for 10 days (C and D). Letters U, H, and S denote unstressed-control, heat stress, and salt stress conditions, respectively.Open in a separate windowFIG. 4.Protein aggregation in Anabaena strains during exposure to heat and salinity stress. The protein aggregation was monitored by ANS fluorescence after 4 h of exposure to 42°C (H) or 10 days of exposure to 150 mM NaCl (S) and compared with the unstressed controls (U) of recombinant strain AnFPNgro and the wild-type Anabaena strain PCC7120 (An7120). The fluorescence intensity output from the spectrofluorimeter is expressed as arbitrary units (a.u.).This study evaluated the possible benefits of groESL overexpression for the general stress tolerance of PCC7120. The recombinant AnFPNgro strain harbored two groESL operons, one native stress-inducible groESL and a second groESL operon integrated at a defined innocuous site and placed downstream of a constitutive PpsbA1 promoter (Fig. (Fig.1).1). The recombinant AnFPNgro strain showed an 8- to 10-fold higher constitutive expression of GroEL under ambient conditions than PCC7120, while its inherent stress-induced GroEL expression was not impaired and resulted in 30- and 48-fold more GroEL under salt and heat stress, respectively (Fig. (Fig.11).The AnFPNgro cells exhibited better growth (Fig. (Fig.2),2), photosynthesis, and nitrogen fixation (Fig. (Fig.3)3) than PCC7120, suggesting a possible limitation on the availability of GroEL under ambient conditions. The protection of photosynthetic pigments and oxygen photoevolution during salinity stress were particularly impressive. Nearly 2- to 2.5-fold higher GroEL levels in AnFPNgro under heat or salt stress, compared to those of PCC7120 (Fig. (Fig.1),1), lowered the stress-triggered protein aggregation (Fig. (Fig.4)4) and had beneficial consequences for photosynthesis and nitrogen fixation in the recombinant strain (Fig. (Fig.3).3). An overall improvement in the aforesaid vital metabolic activities eventually resulted in the superior tolerance of recombinant AnFPNgro to heat and salt stresses.  相似文献   

19.
20.
设为首页 | 免责声明 | 关于勤云 | 加入收藏

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