首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Herpes simplex virus 1 nucleocapsids bud through the inner nuclear membrane (INM) into the perinuclear space to obtain a primary viral envelope. This process requires a protein complex at the INM composed of the UL31 and UL34 gene products. While it is clear that the viral kinase encoded by the US3 gene regulates the localization of pUL31/pUL34 within the INM, the molecular mechanism by which this is accomplished remains enigmatic. Here, we have determined the following. (i) The N terminus of pUL31 is indispensable for the protein''s normal function and contains up to six serines that are phosphorylated by the US3 kinase during infection. (ii) Phosphorylation at these six serines was not essential for a productive infection but was required for optimal viral growth kinetics. (iii) In the presence of active US3 kinase, changing the serines to alanine caused the pUL31/pUL34 complex to aggregate at the nuclear rim and caused some virions to accumulate aberrantly in herniations of the nuclear membrane, much as in cells infected with a US3 kinase-dead mutant. (iv) The replacement of the six serines of pUL31 with glutamic acid largely restored the smooth distribution of pUL34/pUL31 at the nuclear membrane and precluded the accumulation of virions in herniations whether or not US3 kinase was active but also precluded the optimal primary envelopment of nucleocapsids. These observations indicate that the phosphorylation of pUL31 by pUS3 represents an important regulatory event in the virion egress pathway that can account for much of pUS3''s role in nuclear egress. The data also suggest that the dynamics of pUL31 phosphorylation modulate both the primary envelopment and the subsequent fusion of the nascent virion envelope with the outer nuclear membrane.The UL31 and UL34 proteins of herpes simplex virus 1 (HSV-1) form a complex that accumulates at the inner nuclear membrane (INM) of infected cells (26, 27). This complex is essential for the budding of nucleocapsids through the INM into the perinuclear space (26, 28). pUL34 is a type 2 integral membrane protein with a 247-amino-acid nucleoplasmic domain that binds pUL31 and holds the latter in close approximation to the INM (16, 19, 26, 31, 36, 37). Both proteins become incorporated into nascent virions, indicating that they directly or indirectly interact with nucleocapsids during the budding event (27). Interestingly, the coexpression of the pseudorabies virus homologs of HSV pUL31 and pUL34 are sufficient to induce budding from the INM in the absence of other viral proteins (13).The most prominent model of nuclear egress proposes that the step following primary envelopment involves the fusion of the perinuclear virion envelope with the outer nuclear membrane (ONM), allowing subsequent steps in which the deenveloped capsid engages budding sites in the Golgi or trans-Golgi network (20, 32). The US3 protein is a promiscuous kinase that phosphorylates pUL31, pUL34, and several other viral and cellular components (1, 2, 5, 11, 15, 21-23, 25). In the absence of pUS3 kinase activity, (i) virions accumulate within distensions of the perinuclear space that herniate into the nucleoplasm (14, 27, 29), (ii) the pUL31/pUL34 complex is mislocalized at the nuclear rim from a smooth pattern to discrete foci that accumulate adjacent to nuclear membrane herniations (12, 14, 27, 29), and (iii) the onset of infectious virus production is delayed (21, 29).Aberrant accumulations of perinuclear virions similar to those observed in cells infected with US3 kinase-dead viruses have been observed in cells infected with viruses lacking the capacity to produce glycoproteins H and B (gH and gB, respectively) (8). Because these proteins are required for fusion with the plasma membrane or endocytic vesicles during HSV entry (3, 4, 9, 10, 18, 30, 33), it has been proposed that the accumulation of perinuclear virions in the absence of gH and gB reflects a failure in the apparatus that normally mediates the fusion between the nascent virion envelope and the ONM (8). By extension of this hypothesis, pUS3 might act to trigger or otherwise regulate this perinuclear fusion event.The substrate(s) of the pUS3 kinase responsible for the altered localization of the pUL31/pUL34 complex and the aberrant accumulation of perinuclear virions were heretofore unknown. In one study to identify such a substrate, it was determined that precluding the phosphorylation of pUL34 was not responsible for the nuclear egress defects induced by the absence of pUS3 or its kinase activity (29). The current study was therefore undertaken to investigate the hypothesis that the pUS3-mediated phosphorylation of pUL31 is critical to regulate nuclear egress. The presented evidence indicates that aspects of the US3 kinase-dead phenotype, including the retention of virions in the perinuclear space, the mislocalization of the pUL31/pUL34 complex, and the delayed onset of virus replication, can be replicated by precluding pUL31 phosphorylation in the presence or absence of pUS3 kinase activity. The data also suggest that the dynamic phosphorylation of pUL31 is important during the primary envelopment of nucleocapsids.  相似文献   

2.
3.
The UL17 and UL25 proteins (pUL17 and pUL25, respectively) of herpes simplex virus 1 are located at the external surface of capsids and are essential for DNA packaging and DNA retention in the capsid, respectively. The current studies were undertaken to determine whether DNA packaging or capsid assembly affected the pUL17/pUL25 interaction. We found that pUL17 and pUL25 coimmunoprecipitated from cells infected with wild-type virus, whereas the major capsid protein VP5 (encoded by the UL19 gene) did not coimmunoprecipitate with these proteins under stringent conditions. In addition, pUL17 (i) coimmunoprecipitated with pUL25 in the absence of other viral proteins, (ii) coimmunoprecipitated with pUL25 from lysates of infected cells in the presence or absence of VP5, (iii) did not coimmunoprecipitate efficiently with pUL25 in the absence of the triplex protein VP23 (encoded by the UL18 gene), (iv) required pUL25 for proper solubilization and localization within the viral replication compartment, (v) was essential for the sole nuclear localization of pUL25, and (vi) required capsid proteins VP5 and VP23 for nuclear localization and normal levels of immunoreactivity in an indirect immunofluorescence assay. Proper localization of pUL25 in infected cell nuclei required pUL17, pUL32, and the major capsid proteins VP5 and VP23, but not the DNA packaging protein pUL15. The data suggest that VP23 or triplexes augment the pUL17/pUL25 interaction and that VP23 and VP5 induce conformational changes in pUL17 and pUL25, exposing epitopes that are otherwise partially masked in infected cells. These conformational changes can occur in the absence of DNA packaging. The data indicate that the pUL17/pUL25 complex requires multiple viral proteins and functions for proper localization and biochemical behavior in the infected cell.Immature herpes simplex virus (HSV) capsids, like those of all herpesviruses, consist of two protein shells. The outer shell comprises 150 hexons, each composed of six copies of VP5, and 11 pentons, each containing five copies of VP5 (23, 29, 47). One vertex of fivefold symmetry is composed of 12 copies of the protein encoded by the UL6 gene and serves as the portal through which DNA is inserted (22, 39). The pentons and hexons are linked together by 320 triplexes composed of two copies of the UL18 gene product, VP23, and one copy of the UL38 gene product, VP19C (23). Each triplex arrangement has two arms contacting neighboring VP5 subunits (47). The internal shell of the capsid consists primarily of more than 1,200 copies of the scaffold protein ICP35 (VP22a) and a smaller number of protease molecules encoded by the UL26 open reading frame, which self-cleaves to form VP24 and VP21 derived from the amino and carboxyl termini, respectively (11, 12, 19, 25; reviewed in reference 31). The outer shell is virtually identical in the three capsid types found in HSV-infected cells, termed types A, B, and C (5, 6, 7, 29, 43, 48). It is believed that all three are derived from the immature procapsid (21, 38). Type C capsids contain DNA in place of the internal shell, type B capsids contain both shells, and type A capsids consist only of the outer shell (15, 16). Cleavage of viral DNA to produce type C capsids requires not only the portal protein, but all of the major capsid proteins and the products of the UL15, UL17, UL28, UL32, and UL33 genes (2, 4, 10, 18, 26, 28, 35, 46). Only C capsids go on to become infectious virions (27).The outer capsid shell contains minor capsid proteins encoded by the UL25 and UL17 open reading frames (1, 17, 20). These proteins are located on the external surface of the viral capsid (24, 36, 44) and are believed to form a heterodimer arranged as a linear structure, termed the C capsid-specific complex (CCSC), located between pentons and hexons (41). This is consistent with the observation that levels of pUL25 are increased in C capsids as opposed to in B capsids (30). On the other hand, other studies have indicated that at least some UL17 and UL25 proteins (pUL17 and pUL25, respectively) associate with all capsid types, and pUL17 can associate with enveloped light particles, which lack capsid and capsid proteins but contain a number of viral tegument proteins (28, 36, 37). How the UL17 and UL25 proteins attach to capsids is not currently known, although the structure of the CCSC suggests extensive contact with triplexes (41). It is also unclear when pUL17 and pUL25 become incorporated into the capsid during the assembly pathway. Less pUL25 associates with pUL17(−) capsids, suggesting that the two proteins bind capsids either cooperatively or sequentially, although this could also be consequential to the fact that less pUL25 associates with capsids lacking DNA (30, 36).Both pUL25 and pUL17 are necessary for proper nucleocapsid assembly, but their respective deletion generates different phenotypes. Deletion of pUL17 precludes DNA packaging and induces capsid aggregation in the nuclei of infected cells, suggesting a critical early function (28, 34), whereas deletion of pUL25 precludes correct cleavage or retention of full-length cleaved DNA within the capsid (8, 20, 32), thus suggesting a critical function later in the assembly pathway.The current studies were undertaken to determine how pUL17 and pUL25 associate with capsids by studying their interaction and localization in the presence and absence of other capsid proteins.  相似文献   

4.
5.
Incorporation of the herpes simplex virus 1 (HSV-1) portal vertex into the capsid requires interaction with a 12-amino-acid hydrophobic domain within capsid scaffold proteins. The goal of this work was to identify domains and residues in the UL6-encoded portal protein pUL6 critical to the interaction with scaffold proteins. We show that whereas the wild-type portal and scaffold proteins readily coimmunoprecipitated with one another in the absence of other viral proteins, truncation beyond the first 18 or last 36 amino acids of the portal protein precluded this coimmunoprecipitation. The coimmunoprecipitation was also precluded by mutation of conserved tryptophan (W) residues to alanine (A) at positions 27, 90, 127, 163, 241, 262, 532, and 596 of UL6. All of these W-to-A mutations precluded the rescue of a viral deletion mutant lacking UL6, except W163A, which supported replication poorly, and W596A, which fully rescued replication. A recombinant virus bearing the W596A mutation replicated and packaged DNA normally, and scaffold proteins readily coimmunoprecipitated with portal protein from lysates of infected cells. Thus, viral functions compensated for the W596A mutation''s detrimental effects on the portal-scaffold interaction seen during transient expression of portal and scaffold proteins. In contrast, the W27A mutation precluded portal-scaffold interactions in infected cell lysates, reduced the solubility of pUL6, decreased incorporation of the portal into capsids, and abrogated viral-DNA cleavage and packaging.Immature herpesvirus capsids or procapsids consist of two shells: an inner shell, or scaffold, and an outer shell that is roughly spherical and largely composed of the major capsid protein VP5 (24, 38).The capsid scaffold consists of a mixture of the UL26.5 and UL26 gene products, with the UL26.5 gene product (pUL26.5, ICP35, or VP22a) being the most abundant (1, 12, 20, 21, 32, 38). The UL26.5 open reading frame shares its coding frame and C terminus with the UL26 gene but initiates at codon 307 of UL26 (17). The extreme C termini of both VP22a and the UL26-encoded protein (pUL26) interact with the N terminus of VP5 (7, 14, 26, 40, 41). Capsid assembly likely initiates when the portal binds VP5/VP22a and/or VP5/pUL26 complexes (22, 25). The addition of more of these complexes to growing capsid shells eventually produces a closed sphere bearing a single portal. pUL26 within the scaffold contains a protease that cleaves itself between amino acids 247 and 248, separating pUL26 into an N-terminal protease domain called VP24 and a C-terminal domain termed VP21 (4, 5, 8, 9, 28, 42). The protease also cleaves 25 amino acids from pUL26 and VP22a to release VP5 (5, 8, 9). VP21 and VP22a are replaced with DNA when the DNA is packaged (12, 29).When capsids undergo maturation, the outer protein shell angularizes to become icosahedral (13). One fivefold-symmetrical vertex in the angularized outer capsid shell is biochemically distinct from the other 11 and is called the portal vertex because it serves as the channel through which DNA is inserted as it is packaged (23). In herpes simplex virus (HSV), the portal vertex is composed of 12 copies of the portal protein encoded by UL6 (2, 23, 39). We and others have shown that interactions between scaffold and portal proteins are critical for incorporation of the portal into the capsid (15, 33, 44, 45). Twelve amino acids of scaffold proteins are sufficient to interact with the portal protein, and tyrosine and proline resides within this domain are critical for the interaction with scaffold proteins and incorporation of the portal into capsids (45).One goal of the current study was to map domains and residues within the UL6-encoded portal protein that mediate interaction with scaffold proteins. We show that the portal-scaffold interaction requires all but the first 18 and last 36 amino acids of pUL6, as well as several tryptophan residues positioned throughout the portal protein.  相似文献   

6.
7.
Interaction between pUL34 and pUL31 is essential for targeting both proteins to the inner nuclear membrane (INM). Sequences mediating the targeting interaction have been mapped by others with both proteins. We have previously reported identification of charge cluster mutants of herpes simplex virus type 1 UL34 that localize properly to the inner nuclear membrane, indicating interaction with UL31, but fail to complement a UL34 deletion. We have characterized one mutation (CL04) that alters a charge cluster near the N terminus of pUL34 and observed the following. (i) The CL04 mutant has a dominant-negative effect on pUL34 function, indicating disruption of some critical interaction. (ii) In infections with CL04 pUL34, capsids accumulate in close association with the INM, but no perinuclear enveloped viruses, cytoplasmic capsids, or virions or cell surface virions were observed, suggesting that CL04 UL34 does not support INM curvature around the capsid. (iii) Passage of UL34-null virus on a stable cell line that expresses CL04 resulted in selection of extragenic suppressor mutants that grew efficiently using the mutant pUL34. (iv) All extragenic suppressors contained an R229→L mutation in pUL31 that was sufficient to suppress the CL04 phenotype. (v) Immunolocalization and coimmunoprecipitation experiments with truncated forms of pUL34 and pUL31 confirm that N-terminal sequences of pUL34 and a C-terminal domain of pUL31 mediate interaction but not nuclear membrane targeting. pUL34 and pUL31 may make two essential interactions—one for the targeting of the complex to the nuclear envelope and another for nuclear membrane curvature around capsids.Egress of herpesvirus capsids from the nucleus occurs by envelopment of capsids at the inner nuclear membrane (INM) and is followed by de-envelopment at the outer nuclear membrane (ONM). This process can be broken down into a pathway of discrete steps that begin with recruitment of the viral envelopment apparatus to the INM. Herpes simplex virus type 1 (HSV-1) UL34 and UL31 and their homologs in other herpesviruses are required for efficient envelopment at the INM (7, 13, 22, 23, 29). HSV-1 pUL31 and pUL34 are targeted specifically to the INM by a mechanism that requires their interaction with each other (27, 28), and this mutual dependence is a conserved feature of herpesvirus envelopment (9, 14, 27, 28, 32, 33, 39). Localization of these two proteins at the INM results in the recruitment of other proteins, including protein kinase C delta and pUS3, to the nuclear membrane (22, 24, 30). The sequences in HSV-1 pUL34 that mediate interaction with UL31 and that lead to nuclear envelope targeting were mapped to amino acids (aa) 137 to 181 (16). The sequences in the murine cytomegalovirus (MCMV) homolog of UL31, M53, that mediate the nuclear envelope targeting interaction with the UL34 homolog, M50, were mapped to the N-terminal third of the protein in the first of four conserved regions (17), and Schnee et al. subsequently showed that this same region of pUL31 homologs from other families of herpesviruses mediates interaction with the corresponding pUL34 homologs (33).After the targeting of the pUL34/pUL31 complex to the INM, subsequent steps in nuclear egress include, it is thought, (i) local disruption of the nuclear lamina to allow capsid access to the INM, (ii) recognition and docking of capsids by the envelopment apparatus at the INM, (iii) curvature of the inner and outer nuclear membranes around the capsid, (iv) scission of the INM to create an enveloped virion in the space between the INM and ONM, (v) fusion of the virion envelope with the outer nuclear membrane, and (vi) capsid release into the cytoplasm.At least some of the viral and cellular factors critical for nuclear lamina disruption and for de-envelopment fusion have been identified. pUL34, pUL31, and pUS3 of HSV-1 have all been implicated in changes in localization, interaction, and phosphorylation of nuclear lamina components, including lamins A/C and B and the lamina-associated protein, emerin (3, 15, 19, 20, 24, 26, 34, 35). pUS3, pUL31, and glycoproteins B and H have been implicated in de-envelopment of primary virions at the ONM (8, 21, 28, 30, 38).pUL34 and pUL31 are thought to be involved in steps between lamina disruption and de-envelopment, but genetic evidence in infected cells has so far been lacking. Klupp et al. have shown that overexpression of alphaherpesvirus pUL31 and pUL34 in the absence of other viral proteins can induce formation of small vesicles derived from the INM, suggesting a role for these two proteins in membrane curvature around the capsid (12). Tight membrane curvature is an energetically unfavorable event and is thought to be accomplished by coupling curvature to energetically favorable interactions between membrane-bound proteins or protein complexes (reviewed in reference 40). The data of Klupp et al. suggest the possibility that upon recognition of a capsid, pUL31 and pUL34 may interact in a way that induces tight curvature of the INM. Here we present data in support of this hypothesis, showing that a specific point mutation in UL34 induces accumulation of docked capsids at the INM, extragenic suppression of the mutant phenotype is associated with a mutation in UL31, and pUL31 and pUL34 can interact via sequences that are not involved in their INM targeting interaction.We previously published a characterization of a library of 19 charge cluster mutants of pUL34. In each of these mutants, one charge cluster (defined as a group of five consecutive amino acids in which two or more of the residues have charged side chains) was mutated such that the charged residues were replaced by alanine. Six of the 19 charge cluster mutants tested failed to complement replication of UL34-null virus, indicating that they disrupt essential functions of pUL34. Interestingly, five of the six noncomplementing mutants were synthesized at levels comparable to that of wild-type UL34 and localized normally to the nuclear envelope, suggesting that they were unimpaired in their ability to make a nuclear envelope targeting interaction with UL31. In order to identify essential functions of pUL34 downstream of nuclear envelope targeting, we have undertaken a detailed study of the behavior and interactions of these mutants.  相似文献   

8.
9.
Herpes simplex virus type 1 (HSV-1) glycoprotein K (gK) and the UL20 protein (UL20p) are strictly required for virus-induced cell fusion, and mutations within either the gK or UL20 gene cause extensive cell fusion (syncytium formation). We have shown that gK forms a functional protein complex with UL20p, which is required for all gK and UL20p-associated functions in the HSV-1 life cycle. Recently, we showed that the amino-terminal 82 amino acids (aa) of gK (gKa) were required for the expression of the syncytial phenotype of the mutant virus gBΔ28 lacking the carboxyl-terminal 28 amino acids of gB (V. N. Chouljenko, A. V. Iyer, S. Chowdhury, D. V. Chouljenko, and K. G. Kousoulas, J. Virol. 83:12301-12313, 2009). This work suggested that the amino terminus of gK may directly or indirectly interact with gB and/or other viral glycoproteins. Two-way coimmunoprecipitation experiments revealed that UL20p interacted with gB in infected cells. Furthermore, the gKa peptide was coimmunoprecipitated with gB but not gD. Three recombinant baculoviruses were constructed, expressing the amino-terminal 82 aa of gKa together with either the extracellular portion of gB (30 to 748 aa), gD (1 to 340 aa), or gH (1 to 792 aa), respectively. Coimmunoprecipitation experiments revealed that gKa physically interacted with the extracellular portions of gB and gH but not gD. Three additional recombinant baculoviruses expressing gKa and truncated gBs encompassing aa 30 to 154, 30 to 364, and 30 to 500 were constructed. Coimmunoprecipitation experiments showed that gKa physically interacted with all three truncated gBs. Computer-assisted prediction of possible gKa binding sites on gB suggested that gKa may interact predominantly with gB domain I (E. E. Heldwein, H. Lou, F. C. Bender, G. H. Cohen, R. J. Eisenberg, and S. C. Harrison, Science 313:217-220, 2006). These results imply that the gK/UL20p protein complex modulates the fusogenic properties of gB and gH via direct physical interactions.Herpes simplex virus type 1 (HSV-1) can enter into cells via the fusion of its viral envelope with cellular membranes. Also, the virus can spread from infected to uninfected cells by causing virus-induced cell fusion, allowing virions to enter into uninfected cells without being exposed to extracellular spaces. These membrane fusion phenomena are known to be mediated by viral glycoproteins and other viral proteins (reviewed in reference 36). Although wild-type viruses cause a limited amount of virus-induced cell fusion, certain mutations cause extensive virus-induced cell-to-cell fusion (syncytial, or syn, mutations). These syncytial mutations are located predominantly within the UL20 gene (5, 27, 28); the UL24 gene (25, 38); the UL27 gene, encoding glycoprotein gB (7, 15, 18, 32); and the UL53 gene, coding for gK (6, 11, 24, 34, 35, 37).The presence of syncytial mutations within different viral genes, as well as other accumulating evidence, suggests that virus-induced cell fusion is mediated by the concerted action and interactions of the viral glycoproteins gD, gB, and gH/gL as well as gK and the membrane protein UL20p. Specifically, recent studies have shown that gD interacts with both gB and gH/gL (1, 2, 21). However, gB and gH/gL can also interact with each other even in the absence of gD (3). In this membrane fusion model, the binding of gD to its cognate receptors, including nectin-1, herpesvirus entry mediator (HVEM), and other receptors (8, 19, 30, 39-42), is thought to trigger sequential conformational changes in gH/gL and gB causing the fusion of the viral envelope with cellular membranes during virus entry as well as fusion among cellular membranes (22, 23). The transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (31, 43), suggesting that these four viral glycoproteins are necessary and sufficient for membrane fusion. However, this transient fusion system does not accurately depict virus-induced cell fusion. Specifically, viral glycoprotein K (gK) and the UL20 membrane protein (UL20p) have been shown to be strictly required for virus-induced cell fusion (10, 27, 29). Moreover, syncytial mutations within gK (6, 11, 24, 34, 35, 37) or UL20 (5, 27, 28) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than the wild-type virus into susceptible cells (17). In contrast, the transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while the coexpression of wild-type gK with gB, gD, and gH/gL was reported previously to inhibit cell fusion in certain cell lines (4). To date, there is no direct evidence that either gK or UL20p interacts with gB, gD, gH, or gL.The X-ray structure of the ectodomain of HSV-1 gB has been determined and was predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB (23). Single-amino-acid changes within the carboxyl terminus of gB located intracellularly as well as the deletion of the terminal 28 amino acids (aa) of gB cause extensive virus-induced cell fusion, presumably because they alter the extracellular conformation of gB (15, 31, 43). We have previously shown that HSV-1 gK and UL20p functionally and physically interact and that these interactions are absolutely necessary for their coordinate intracellular transport, cell surface expression, and functions in the HSV-1 life cycle (13, 16). In contrast to gB, syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (12), while syncytial mutations of UL20 are located within the amino terminus of UL20p shown to be located intracellularly (27).Recently, we showed that the a peptide composed of the amino-terminal 82 amino acids of gK (gKa) can complement in trans for gB-mediated cell fusion caused by the deletion of the carboxyl-terminal 28 amino acids of gB, suggesting that the gKa peptide interacted with gB or other viral glycoproteins involved in virus-induced cell fusion (10). In this work, we demonstrate that UL20p and the amino terminus of gKa physically interact with gB in infected cells, while the gKa peptide is also capable of binding to the extracellular portion of gH, suggesting that gK/UL20p modulates virus-induced cell fusion via direct interactions with gB and gH.  相似文献   

10.
Human cytomegalovirus (HCMV) UL37 proteins traffic sequentially from the endoplasmic reticulum (ER) to the mitochondria. In transiently transfected cells, UL37 proteins traffic into the mitochondrion-associated membranes (MAM), the site of contact between the ER and mitochondria. In HCMV-infected cells, the predominant UL37 exon 1 protein, pUL37x1, trafficked into the ER, the MAM, and the mitochondria. Surprisingly, a component of the MAM calcium signaling junction complex, cytosolic Grp75, was increasingly enriched in heavy MAM from HCMV-infected cells. These studies show the first documented case of a herpesvirus protein, HCMV pUL37x1, trafficking into the MAM during permissive infection and HCMV-induced alteration of the MAM protein composition.The human cytomegalovirus (HCMV) UL37 immediate early (IE) locus expresses multiple products, including the predominant UL37 exon 1 protein, pUL37x1, also known as viral mitochondrion-localized inhibitor of apoptosis (vMIA), during lytic infection (16, 22, 24, 39, 44). The UL37 glycoprotein (gpUL37) shares UL37x1 sequences and is internally cleaved, generating pUL37NH2 and gpUL37COOH (2, 22, 25, 26). pUL37x1 is essential for the growth of HCMV in humans (17) and for the growth of primary HCMV strains (20) and strain AD169 (14, 35, 39, 49) but not strain TownevarATCC in permissive human fibroblasts (HFFs) (27).pUL37x1 induces calcium (Ca2+) efflux from the endoplasmic reticulum (ER) (39), regulates viral early gene expression (5, 10), disrupts F-actin (34, 39), recruits and inactivates Bax at the mitochondrial outer membrane (MOM) (4, 31-33), and inhibits mitochondrial serine protease at late times of infection (28).Intriguingly, HCMV UL37 proteins localize dually in the ER and in the mitochondria (2, 9, 16, 17, 24-26). In contrast to other characterized, similarly localized proteins (3, 6, 11, 23, 30, 38), dual-trafficking UL37 proteins are noncompetitive and sequential, as an uncleaved gpUL37 mutant protein is ER translocated, N-glycosylated, and then imported into the mitochondria (24, 26).Ninety-nine percent of ∼1,000 mitochondrial proteins are synthesized in the cytosol and directly imported into the mitochondria (13). However, the mitochondrial import of ER-synthesized proteins is poorly understood. One potential pathway is the use of the mitochondrion-associated membrane (MAM) as a transfer waypoint. The MAM is a specialized ER subdomain enriched in lipid-synthetic enzymes, lipid-associated proteins, such as sigma-1 receptor, and chaperones (18, 45). The MAM, the site of contact between the ER and the mitochondria, permits the translocation of membrane-bound lipids, including ceramide, between the two organelles (40). The MAM also provides enriched Ca2+ microdomains for mitochondrial signaling (15, 36, 37, 43, 48). One macromolecular MAM complex involved in efficient ER-to-mitochondrion Ca2+ transfer is comprised of ER-bound inositol 1,4,5-triphosphate receptor 3 (IP3R3), cytosolic Grp75, and a MOM-localized voltage-dependent anion channel (VDAC) (42). Another MAM-stabilizing protein complex utilizes mitofusin 2 (Mfn2) to tether ER and mitochondrial organelles together (12).HCMV UL37 proteins traffic into the MAM of transiently transfected HFFs and HeLa cells, directed by their NH2-terminal leaders (8, 47). To determine whether the MAM is targeted by UL37 proteins during infection, we fractionated HCMV-infected cells and examined pUL37x1 trafficking in microsomes, mitochondria, and the MAM throughout all temporal phases of infection. Because MAM domains physically bridge two organelles, multiple markers were employed to verify the purity and identity of the fractions (7, 8, 19, 46, 47).(These studies were performed in part by Chad Williamson in partial fulfillment of his doctoral studies in the Biochemistry and Molecular Genetics Program at George Washington Institute of Biomedical Sciences.)HFFs and life-extended (LE)-HFFs were grown and not infected or infected with HCMV (strain AD169) at a multiplicity of 3 PFU/cell as previously described (8, 26, 47). Heavy (6,300 × g) and light (100,000 × g) MAM fractions, mitochondria, and microsomes were isolated at various times of infection and quantified as described previously (7, 8, 47). Ten- or 20-μg amounts of total lysate or of subcellular fractions were resolved by SDS-PAGE in 4 to 12% Bis-Tris NuPage gels (Invitrogen) and examined by Western analyses (7, 8, 26). Twenty-microgram amounts of the fractions were not treated or treated with proteinase K (3 μg) for 20 min on ice, resolved by SDS-PAGE, and probed by Western analysis. The blots were probed with rabbit anti-UL37x1 antiserum (DC35), goat anti-dolichyl phosphate mannose synthase 1 (DPM1), goat anti-COX2 (both from Santa Cruz Biotechnology), mouse anti-Grp75 (StressGen Biotechnologies), and the corresponding horseradish peroxidase-conjugated secondary antibodies (8, 47). Reactive proteins were detected by enhanced chemiluminescence (ECL) reagents (Pierce), and images were digitized as described previously (26, 47).  相似文献   

11.
12.
Human cytomegalovirus (HCMV) is a widely circulating pathogen that causes severe disease in immunocompromised patients and infected fetuses. By immortalizing memory B cells from HCMV-immune donors, we isolated a panel of human monoclonal antibodies that neutralized at extremely low concentrations (90% inhibitory concentration [IC90] values ranging from 5 to 200 pM) HCMV infection of endothelial, epithelial, and myeloid cells. With the single exception of an antibody that bound to a conserved epitope in the UL128 gene product, all other antibodies bound to conformational epitopes that required expression of two or more proteins of the gH/gL/UL128-131A complex. Antibodies against gB, gH, or gM/gN were also isolated and, albeit less potent, were able to neutralize infection of both endothelial-epithelial cells and fibroblasts. This study describes unusually potent neutralizing antibodies against HCMV that might be used for passive immunotherapy and identifies, through the use of such antibodies, novel antigenic targets in HCMV for the design of immunogens capable of eliciting previously unknown neutralizing antibody responses.Human cytomegalovirus (HCMV) is a member of the herpesvirus family which is widely distributed in the human population and can cause severe disease in immunocompromised patients and upon infection of the fetus. HCMV infection causes clinical disease in 75% of patients in the first year after transplantation (58), while primary maternal infection is a major cause of congenital birth defects including hearing loss and mental retardation (5, 33, 45). Because of the danger posed by this virus, development of an effective vaccine is considered of highest priority (51).HCMV infection requires initial interaction with the cell surface through binding to heparan sulfate proteoglycans (8) and possibly other surface receptors (12, 23, 64, 65). The virus displays a broad host cell range (24, 53), being able to infect several cell types such as endothelial cells, epithelial cells (including retinal cells), smooth muscle cells, fibroblasts, leukocytes, and dendritic cells (21, 37, 44, 54). Endothelial cell tropism has been regarded as a potential virulence factor that might influence the clinical course of infection (16, 55), whereas infection of leukocytes has been considered a mechanism of viral spread (17, 43, 44). Extensive propagation of HCMV laboratory strains in fibroblasts results in deletions or mutations of genes in the UL131A-128 locus (1, 18, 21, 36, 62, 63), which are associated with the loss of the ability to infect endothelial cells, epithelial cells, and leukocytes (15, 43, 55, 61). Consistent with this notion, mouse monoclonal antibodies (MAbs) to UL128 or UL130 block infection of epithelial and endothelial cells but not of fibroblasts (63). Recently, it has been shown that UL128, UL130, and UL131A assemble with gH and gL to form a five-protein complex (thereafter designated gH/gL/UL128-131A) that is an alternative to the previously described gCIII complex made of gH, gL, and gO (22, 28, 48, 63).In immunocompetent individuals T-cell and antibody responses efficiently control HCMV infection and reduce pathological consequences of maternal-fetal transmission (13, 67), although this is usually not sufficient to eradicate the virus. Albeit with controversial results, HCMV immunoglobulins (Igs) have been administered to transplant patients in association with immunosuppressive treatments for prophylaxis of HCMV disease (56, 57), and a recent report suggests that they may be effective in controlling congenital infection and preventing disease in newborns (32). These products are plasma derivatives with relatively low potency in vitro (46) and have to be administered by intravenous infusion at very high doses in order to deliver sufficient amounts of neutralizing antibodies (4, 9, 32, 56, 57, 66).The whole spectrum of antigens targeted by HCMV-neutralizing antibodies remains poorly characterized. Using specific immunoabsorption to recombinant antigens and neutralization assays using fibroblasts as model target cells, it was estimated that 40 to 70% of the serum neutralizing activity is directed against gB (6). Other studies described human neutralizing antibodies specific for gB, gH, or gM/gN viral glycoproteins (6, 14, 26, 29, 34, 41, 52, 60). Remarkably, we have recently shown that human sera exhibit a more-than-100-fold-higher potency in neutralizing infection of endothelial cells than infection of fibroblasts (20). Similarly, CMV hyperimmunoglobulins have on average 48-fold-higher neutralizing activities against epithelial cell entry than against fibroblast entry (10). However, epitopes that are targeted by the antibodies that comprise epithelial or endothelial cell-specific neutralizing activity of human immune sera remain unknown.In this study we report the isolation of a large panel of human monoclonal antibodies with extraordinarily high potency in neutralizing HCMV infection of endothelial and epithelial cells and myeloid cells. With the exception of a single antibody that recognized a conserved epitope of UL128, all other antibodies recognized conformational epitopes that required expression of two or more proteins of the gH/gL/UL128-131A complex.  相似文献   

13.
Of the four required herpes simplex virus (HSV) entry glycoproteins, the precise role of gH-gL in fusion remains the most elusive. The heterodimer gH-gL has been proposed to mediate hemifusion after the interaction of another required glycoprotein, gD, with a receptor. To identify functional domains of HSV-1 gH, we generated 22 randomized linker-insertion mutants. Analyses of 22 gH mutants revealed that gH is relatively tolerant of insertion mutations, as 15 of 22 mutants permitted normal processing and transport of gH-gL to the cell surface. gH mutants that were not expressed well at the cell surface did not function in fusion or viral entry. The screening of gH mutants for function revealed the following: (i) for wild-type gH and some gH mutants, fusion with nectin-1-expressing target cells occurred more rapidly than with herpesvirus entry mediator (HVEM)-expressing target cells; (ii) some gH mutants reduced the rate of cell fusion without abrogating fusion completely, indicating that gH may play a role in governing the kinetics of fusion and may be responsible for a rate-limiting first stage in HSV-1 fusion; and (iii) only one gH mutant, located within the short cytoplasmic tail, completely abrogated function, indicating that the gH cytoplasmic tail is crucial for cell fusion and viral infectivity.Herpes simplex virus (HSV), an enveloped neurotropic virus, infects target cells via membrane fusion, a process executed by viral fusion proteins capable of inserting into target membranes. Unlike many enveloped viruses that induce fusion through the activity of a single viral fusion protein, HSV requires four glycoproteins, glycoprotein B (gB), glycoprotein D (gD), glycoprotein H (gH), and glycoprotein L (gL), to execute fusion (6, 40, 42). The focus of this study, gH, is expressed as a heterodimer with gL (gH-gL). HSV gH and gL rely on one another for proper folding, posttranslational processing, and transport to the cell and virion surface (5, 23, 35).A sequential model of entry is the prevailing working hypothesis of HSV entry (1-3, 28, 32, 41). Viral attachment is mediated by the binding of glycoprotein C (gC) or gB to cell surface glycosaminoglycans such as heparan sulfate (38). The subsequent fusion between the virion envelope and host cell membrane is thought to result from a series of concerted events. First, gD binds to one of its host cell receptors. These receptors include herpesvirus entry mediator (HVEM), a member of the tumor necrosis factor (TNF) receptor family; nectin-1 and nectin-2, cell adhesion molecules of the Ig superfamily; and heparan sulfate modified by specific 3-O-sulfotransferases (39).It was previously proposed that gD binding a receptor induces a conformational change that allows for interactions between gD, gB, and/or gH-gL (1, 2, 8, 10, 16, 25, 32). It is thought that while gD functions primarily in receptor binding, gB and gH-gL function as the core fusion machinery of HSV.Based on its crystal structure, gB has structural features typical of viral fusion proteins in general and is structurally similar to vesicular stomatitis virus (VSV) glycoprotein G, the fusion protein of VSV (22, 34). In addition to its resemblance to other viral fusogens, gB also binds its own receptor, paired immunoglobulin-like receptor (PILRalpha) (36, 37). Importantly, HSV gB does not successfully execute fusion in the absence of gD or gH-gL (41). Compared to the other required HSV entry glycoproteins, relatively little is known about the specific roles of gH-gL during fusion. The structure of gH-gL is unknown, although in silico analyses and studies of synthetic gH peptides suggested that gH also has fusogenic properties (12, 13, 17-20).gD, a gD receptor, and gH-gL have been shown to be sufficient for inducing hemifusion, the mixing of the proximal leaflets of the viral and host cell bilayers (41). Several lines of research suggest that the subsequent step in fusion is an interaction between gH-gL and gB, with the latter glycoprotein being required for a committed and expanding fusion pore (1-3, 16, 28, 41). However, it is still unclear whether the gB and gH-gL interaction requires that gD first bind a receptor (1, 3), indicating that another viable model of HSV entry may be nonsequential gD-gB-gH-gL complex formation.Several domains important for fusion within HSV gH have been discerned. The only function associated with the N-terminal domain of HSV gH, to date, is gL binding. Residues 377 to 397 within a predicted alpha-helix in the gH ectodomain are required for cell-cell fusion and complementation of a gH-null virus (18). The mutation of a predicted heptad repeat region spanning residues 443 to 471 abrogated cell-cell fusion (17). Insertion mutations within what has been termed the pretransmembrane region of gH have also been shown to abrogate fusion and viral entry (11). The glycine residue at position 812 within the predicted gH transmembrane domain was shown previously to be important for fusion (21). Finally, although the deletion of the final six residues of gH (residues 832 to 838), which are within its short cytoplasmic tail, has no effect on fusion, further deletions were shown to decrease polykaryocyte formation by a syncytial HSV strain (4, 43).We used a transposon-based comprehensive random linker-insertion mutagenesis strategy to generate a library of mutants spanning the entire length of HSV-1 gH, an 838-amino-acid type I membrane protein. A panel of 22 insertion mutants was generated, 15 of which were expressed at near-normal levels on the cell surface. Interestingly, some insertions reduced the rate of cell fusion rather than abrogating cell fusion activity altogether, suggesting that gH may have a role in governing the kinetics of fusion and may be responsible for a rate-limiting first stage in HSV-1 fusion. Additionally, one insertion mutation that completely abrogated cell fusion and viral infectivity is located within the gH cytoplasmic tail, indicating that the short C-terminal tail of gH is critical for cell fusion and entry mediated by HSV-1.  相似文献   

14.
Bovine herpesvirus 1 (BoHV-1) and BoHV-5 are closely related pathogens of cattle, but only BoHV-5 is considered a neuropathogen. We engineered intertypic gD exchange mutants with BoHV-1 and BoHV-5 backbones in order to address their in vitro and in vivo host ranges, with particular interest in invasion of the brain. The new viruses replicated in cell culture with similar dynamics and to titers comparable to those of their wild-type parents. However, gD of BoHV-5 (gD5) was able to interact with a surprisingly broad range of nectins. In vivo, gD5 provided a virulent phenotype to BoHV-1 in AR129 mice, featuring a high incidence of neurological symptoms and early onset of disease. However, only virus with the BoHV-5 backbone, independent of the gD type, was detected in the brain by immunohistology. Thus, gD of BoHV-5 confers an extended cellular host range to BoHV-1 and may be considered a virulence factor but does not contribute to the invasion of the brain.Bovine herpesvirus 1 (BoHV-1) and BoHV-5 belong to the subfamily Alphaherpesvirinae and are closely related pathogens of cattle (22). The protein repertoire of the two viruses averages 82% amino acid identity (20). Both viruses are neurotropic, but only BoHV-5 can significantly replicate in the central nervous system (CNS) to cause encephalitis of either naturally infected cattle or experimentally inoculated laboratory animals (2, 5, 6, 12, 40, 41, 44). Glycoprotein D (gD) is accepted as the critical and essential receptor-binding protein of many alphaherpesviruses (reviewed in references 8 and 48). The main gD receptors identified to date include members of the tumor necrosis factor (TNF) receptor family (HveA) and the poliovirus receptor family (HveB or nectin 2 and HveC or nectin 1) (28, 42, 51). Furthermore, a modified form of heparan sulfate, 3-O-sulfated heparan sulfate, can mediate herpesvirus entry (46). J1.1-2 cells (J cells) represent a subpopulation of thymidine kinase-negative baby hamster kidney (BHK) cells selected for their property of being resistant to infection with herpes simplex virus type 1 (HSV-1), HSV-2, and BoHV-1. The expression of nectin 1 in those cells rendered them susceptible to BoHV-1 infection and replication, which suggests that nectin 1 can serve as a receptor for BoHV-1 gD (gD1) (16, 18, 28). Interestingly, we observed that BoHV-5 was able to productively replicate in J cells without the nectin 1 receptor.According to a previously reported sequence comparison of BoHV-1 and BoHV-5 (20), the highest divergence between the two viruses mapped to the latency-related region and the immediate-early proteins (less than 75% amino acid identity) BICP0, BICP4, and BICP22. Glycoprotein E (gE) was also listed in this category, with 74% amino acid identity between gE of BoHV-1 (gE1) and gE5. This fact also gave ample reason for attempts to map the neurovirulent phenotype of BoHV-5 to the gE5 molecule (3, 4, 13). In contrast, the highest sequence similarities between the two viruses were described for proteins involved in viral DNA replication and processing as well as certain virion proteins. Among others, the predicted amino acid sequences of gD1 and gD5 were listed as being 98% identical (20). However, our own analysis using the European Molecular Biology software suite (43) revealed only 79.9% amino acid identity. Obviously, the most extensive difference between gD1 and gD5 maps to a glycine-rich stretch located in the molecule''s ectodomain, between amino acids (aa) 280 and 330 of gD5, in close vicinity to the transmembrane region.Based on these considerations, we hypothesized that BoHV-5 was able to make use of a cellular receptor that is unavailable to BoHV-1. To test this hypothesis, the gD genes were removed from bacterial artificial chromosomes (BACs) harboring the genome of either BoHV-5 or BoHV-1 (27). In a second step, gD exchange viruses were created by the cotransfection of the gD-less BACs with appropriate plasmids carrying either the gD1 or gD5 gene and appropriate flanking sequences. The newly generated viruses included an intertypic BoHV-5 mutant carrying gD1 in the place of gD5 and a corresponding BoHV-1 carrying gD5. These mutants, together with appropriate revertant mutants, were then used to explore their ability to infect J cells in vitro and their ability to cause neurological disease and invade the brain in vivo, in a previously established mouse model (2). Our results indicate that gD5 confers an extended host range to BoHV-1 but is nonessential for the invasion of the brain.  相似文献   

15.
16.
Herpes simplex virus type 1 (HSV-1)-induced cell fusion is mediated by viral glycoproteins and other membrane proteins expressed on infected cell surfaces. Certain mutations in the carboxyl terminus of HSV-1 glycoprotein B (gB) and in the amino terminus of gK cause extensive virus-induced cell fusion. Although gB is known to be a fusogenic glycoprotein, the mechanism by which gK is involved in virus-induced cell fusion remains elusive. To delineate the amino-terminal domains of gK involved in virus-induced cell fusion, the recombinant viruses gKΔ31-47, gKΔ31-68, and gKΔ31-117, expressing gK carrying in-frame deletions spanning the amino terminus of gK immediately after the gK signal sequence (amino acids [aa] 1 to 30), were constructed. Mutant viruses gKΔ31-47 and gKΔ31-117 exhibited a gK-null (ΔgK) phenotype characterized by the formation of very small viral plaques and up to a 2-log reduction in the production of infectious virus in comparison to that for the parental HSV-1(F) wild-type virus. The gKΔ31-68 mutant virus formed substantially larger plaques and produced 1-log-higher titers than the gKΔ31-47 and gKΔ31-117 mutant virions at low multiplicities of infection. Deletion of 28 aa from the carboxyl terminus of gB (gBΔ28syn) caused extensive virus-induced cell fusion. However, the gBΔ28syn mutation was unable to cause virus-induced cell fusion in the presence of the gKΔ31-68 mutation. Transient expression of a peptide composed of the amino-terminal 82 aa of gK (gKa) produced a glycosylated peptide that was efficiently expressed on cell surfaces only after infection with the HSV-1(F), gKΔ31-68, ΔgK, or UL20-null virus. The gKa peptide complemented the gKΔ31-47 and gKΔ31-68 mutant viruses for infectious-virus production and for gKΔ31-68/gBΔ28syn-mediated cell fusion. These data show that the amino terminus of gK modulates gB-mediated virus-induced cell fusion and virion egress.Herpes simplex virus type 1 (HSV-1) specifies at least 11 virally encoded glycoproteins, as well as several nonglycosylated and lipid-anchored membrane-associated proteins, which serve important functions in virion infectivity and virus spread. Although cell-free enveloped virions can efficiently spread viral infection, virions can also spread by causing cell fusion of adjacent cellular membranes. Virus-induced cell fusion, which is caused by viral glycoproteins expressed on infected cell surfaces, enables transmission of virions from one cell to another, avoiding extracellular spaces and exposure of free virions to neutralizing antibodies (reviewed in reference 56). Most mutations that cause extensive virus-induced cell-to-cell fusion (syncytial or syn mutations) have been mapped to at least four regions of the viral genome: the UL20 gene (5, 42, 44); the UL24 gene (37, 58); the UL27 gene, encoding glycoprotein B (gB) (9, 51); and the UL53 gene, coding for gK (7, 15, 35, 53, 54, 57).Increasing evidence suggests that virus-induced cell fusion is mediated by the concerted action of glycoproteins gD, gB, and gH/gL. Recent studies have shown that gD interacts with both gB and gH/gL (1, 2). Binding of gD to its cognate receptors, including Nectin-1, HVEM, and others (12, 29, 48, 59, 60, 62, 63), is thought to trigger conformation changes in gH/gL and gB that cause fusion of the viral envelope with cellular membranes during virus entry and virus-induced cell fusion (32, 34). Transient coexpression of gB, gD, and gH/gL causes cell-to-cell fusion (49, 68). However, this phenomenon does not accurately model viral fusion, because other viral glycoproteins and membrane proteins known to be important for virus-induced cell fusion are not required (6, 14, 31). Specifically, gK and UL20 were shown to be absolutely required for virus-induced cell fusion (21, 46). Moreover, syncytial mutations within gK (7, 15, 35, 53, 54, 57) or UL20 (5, 42, 44) promote extensive virus-induced cell fusion, and viruses lacking gK enter more slowly than wild-type virus into susceptible cells (25). Furthermore, transient coexpression of gK carrying a syncytial mutation with gB, gD, and gH/gL did not enhance cell fusion, while coexpression of the wild-type gK with gB, gD, and gH/gL inhibited cell fusion (3).Glycoproteins gB and gH are highly conserved across all subfamilies of herpesviruses. gB forms a homotrimeric type I integral membrane protein, which is N glycosylated at multiple sites within the polypeptide. An unusual feature of gB is that syncytial mutations that enhance virus-induced cell fusion are located exclusively in the carboxyl terminus of gB, which is predicted to be located intracellularly (51). Single-amino-acid substitutions within two regions of the intracellular cytoplasmic domain of gB were shown to cause syncytium formation and were designated region I (amino acid [aa] positions 816 and 817) and region II (aa positions 853, 854, and 857) (9, 10, 28, 69). Furthermore, deletion of 28 aa from the carboxyl terminus of gB, disrupting the small predicted alpha-helical domain H17b, causes extensive virus-induced cell fusion as well as extensive glycoprotein-mediated cell fusion in the gB, gD, and gH/gL transient-coexpression system (22, 49, 68). The X-ray structure of the ectodomain of gB has been determined and is predicted to assume at least two major conformations, one of which may be necessary for the fusogenic properties of gB. Therefore, perturbation of the carboxyl terminus of gB may alter the conformation of the amino terminus of gB, thus favoring one of the two predicted conformational structures that causes membrane fusion (34).The UL53 (gK) and UL20 genes encode multipass transmembrane proteins of 338 and 222 aa, respectively, which are conserved in all alphaherpesviruses (15, 42, 55). Both proteins have multiple sites where posttranslational modification can occur; however, only gK is posttranslationally modified by N-linked carbohydrate addition (15, 35, 55). The specific membrane topologies of both gK and UL20 protein (UL20p) have been predicted and experimentally confirmed using epitope tags inserted within predicted intracellular and extracellular domains (18, 21, 44). Syncytial mutations in gK map predominantly within extracellular domains of gK and particularly within the amino-terminal portion of gK (domain I) (18), while syncytial mutations of UL20 are located within the amino terminus of UL20p, shown to be located intracellularly (44). A series of recent studies have shown that HSV-1 gK and UL20 functionally and physically interact and that these interactions are necessary for their coordinate intracellular transport and cell surface expression (16, 18, 21, 26, 45). Specifically, direct protein-protein interactions between the amino terminus of HSV-1 UL20 and gK domain III, both of which are localized intracellularly, were recently demonstrated by two-way coimmunoprecipitation experiments (19).According to the most prevalent model for herpesvirus intracellular morphogenesis, capsids initially assemble within the nuclei and acquire a primary envelope by budding into the perinuclear spaces. Subsequently, these virions lose their envelope through fusion with the outer nuclear lamellae. Within the cytoplasm, tegument proteins associate with the viral nucleocapsid and final envelopment occurs by budding of cytoplasmic capsids into specific trans-Golgi network (TGN)-associated membranes (8, 30, 47, 70). Mature virions traffic to cell surfaces, presumably following the cellular secretory pathway (33, 47, 61). In addition to their significant roles in virus-induced cell fusion, gK and UL20 are required for cytoplasmic virion envelopment. Viruses with deletions in either the gK or the UL20 gene are unable to translocate from the cytoplasm to extracellular spaces and accumulated as unenveloped virions in the cytoplasm (5, 15, 20, 21, 26, 35, 36, 38, 44, 55). Current evidence suggests that the functions of gK and UL20 in cytoplasmic virion envelopment and virus-induced cell fusion are carried out by different, genetically separable domains of UL20p. Specifically, UL20 mutations within the amino and carboxyl termini of UL20p allowed cotransport of gK and UL20p to cell surfaces, virus-induced cell fusion, and TGN localization, while effectively inhibiting cytoplasmic virion envelopment (44, 45).In this paper, we demonstrate that the amino terminus of gK expressed as a free peptide of 82 aa (gKa) is transported to infected cell surfaces by viral proteins other than gK or UL20p and facilitates virus-induced cell fusion caused by syncytial mutations in the carboxyl terminus of gB. Thus, functional domains of gK can be genetically separated, as we have shown previously (44, 45), as well as physically separated into different peptide portions that retain functional activities of gK. These results are consistent with the hypothesis that the amino terminus of gK directly or indirectly interacts with and modulates the fusogenic properties of gB.  相似文献   

17.
Infection of quiescent cells by human cytomegalovirus (HCMV) elicits severe cell cycle deregulation, resulting in a G1/S arrest, which can be partly attributed to the inactivation of the anaphase-promoting complex (APC). As we previously reported, the premature phosphorylation of its coactivator Cdh1 and/or the dissociation of the core complex can account for the inactivation. We have expanded on these results and further delineated the key components required for disabling the APC during HCMV infection. The viral protein kinase UL97 was hypothesized to phosphorylate Cdh1, and consistent with this, phosphatase assays utilizing a virus with a UL97 deletion mutation (ΔUL97 virus) indicated that Cdh1 is hypophosphorylated at early times in the infection. Mass spectrometry analysis demonstrated that UL97 can phosphorylate Cdh1 in vitro, and the majority of the sites identified correlated with previously characterized cyclin-dependent kinase (Cdk) consensus sites. Analysis of the APC core complex during ΔUL97 virus infection showed APC dissociation occurring at the same time as during infection with wild-type virus, suggesting that the UL97-mediated phosphorylation of Cdh1 is not required for this to occur. Further investigation of the APC subunits showed a proteasome-dependent loss of the APC5 and APC4 subunits that was temporally associated with the disassembly of the APC. Immediate early viral gene expression was not sufficient for the degradation of APC4 and APC5, indicating that a viral early gene product(s), possibly in association with a de novo-synthesized cellular protein(s), is involved.Human cytomegalovirus (HCMV), a highly prevalent β-herpesvirus, can cause serious birth defects and disease in immunocompromised individuals, and it may be associated with cancer and cardiovascular disease (53). Viral gene expression is temporally regulated and is dependent on many cellular factors for a productive infection. Immediate early (IE) genes are expressed by 2 h postinfection (p.i.) and transactivate the early genes required for viral DNA replication. The expression of the late genes, which encode proteins involved in virion maturation and egress, is dependent on viral DNA replication.The virus has adopted different strategies for altering the cellular environment to make it more conducive to productive infection, including the stimulation of host cell DNA replication pathways, cell cycle deregulation and arrest, immune evasion, and inhibition of apoptosis (53). Although HCMV encodes its own DNA polymerase, it is dependent on other cellular resources for DNA replication. Infection of quiescent cells induces passage toward S phase such that the host cell is stimulated to generate proteins and DNA precursors necessary for genome replication; however, entry into S phase and cellular DNA replication are subsequently blocked and the cell arrests in G1/S (1, 10, 11, 14, 30, 45). Cellular resources are thereby presumably free to be efficiently utilized for viral replication. Cell cycle arrest by HCMV is achieved in part through the misregulation of several cell cycle proteins, including the phosphorylation and accumulation of the Rb family pocket proteins, upregulation of cyclins E and B and their associated kinase activities, inhibition of cyclin A expression, stabilization of p53, and accumulation of Cdc6 and geminin, which inhibits licensing of the cellular origins of DNA replication (8, 17, 30, 49, 54, 65). Some of these cell cycle defects can be attributed to a deregulation of the anaphase-promoting complex (APC) (8, 72, 79, 80), an E3 ubiquitin ligase that is responsible for the timely degradation of cell cycle proteins and mitotic cyclins to promote cycle progression from mitosis through G1 to S phase (58, 74). As the APC also appears to be a common target among other viruses, including the chicken anemia virus, adenoviruses, and poxviruses (23, 36, 52, 70), understanding the mechanisms leading to its inactivation during viral infection has been of great interest.As we have previously reported, multiple mechanisms may be involved in disabling the APC during HCMV infection (72), which is not surprising given the complexity of its structure and regulation (for a review, see references 58 and 74). The APC is a large multisubunit complex consisting of at least 11 conserved core subunits, as well as other species-specific subunits. In metazoans, the APC2 and APC11 subunits form the catalytic core, and along with APC10, provide the platform for binding the E2 ubiquitin-conjugating enzyme. Each of the APC3, APC8, APC6, and APC7 subunits contain multiple copies of the tetratricopeptide repeat (TPR) motif and together make up the TPR subcomplex, which provides a platform of protein interaction surfaces for binding the coactivators (i.e., Cdh1 and Cdc20) and various substrates. These two subcomplexes are bridged by the large scaffolding subunit APC1, with the TPR subcomplex tethered to APC1 through APC4 and APC5. The binding between APC1, APC4, APC5, and APC8 is also interdependent, such that the loss of one subunit decreases the association of the other three (71).The APC is activated by either of its coactivators, Cdh1 or Cdc20, which also function in recruiting specific substrates to the APC during different phases of the cell cycle. The phosphorylation of several APC subunits at the onset of mitosis, including APC1 and the TPR subunits, by cyclin B/cyclin-dependent kinase 1 (Cdk1) and Plk1 allows the binding of Cdc20 and subsequent activation of the APC (APCCdc20) (19, 37), whereas the binding and activation of the complex by Cdh1 is inhibited through its phosphorylation by cyclin B/Cdk1 (9, 29, 38, 83). As cells pass the spindle assembly checkpoint, APCCdc20 ubiquitinates securin (to allow for sister chromatid separation) and cyclin B for degradation by the proteasome (42, 67). The subsequent inactivation of Cdk1 and activation of mitotic phosphatases during late anaphase relieves the inhibitory phosphorylation on Cdh1, presumably by Cdc14 (6, 38, 44), which then allows Cdh1 to bind and activate the APC (APCCdh1). APCCdh1 ubiquitinates Cdc20 and mitotic cyclins for degradation to facilitate mitotic exit and maintains their low levels, along with S-phase regulators (e.g., Cdc6, geminin, etc.), during G1 (16, 50, 59, 63). The inactivation of APCCdh1 as cells enter S phase may be mediated in part through the phosphorylation of Cdh1 by cyclin A/Cdk2 (46) and Cdh1 binding to the inhibitor Emi1 (25). The inactivation of Cdh1 by phosphorylation has been shown in all organisms studied thus far (e.g., yeast, Drosophila, plants, mammals, etc.), and mutants mimicking constitutively phosphorylated Cdh1 on Cdk consensus sites can neither bind nor activate the APC in vivo or in vitro (9, 29, 38, 69, 83).During HCMV infection of fibroblasts in G0/G1, however, Cdh1 becomes prematurely phosphorylated in a Cdk-independent manner and no longer associates with the APC (72). This dissociation does not appear to be due to an overexpression of Emi1 (79). Cdc20 also can no longer associate with the APC (79), suggesting a defect in the APC core. We have further shown that the APC core complex disassembles during the infection, with the TPR subunits (i.e., APC3, APC7, and APC8) and APC10 localizing to the cytosol, while APC1 remains nuclear (72). Interestingly, both the phosphorylation of Cdh1 and the dissociation of the APC occur at similar times during HCMV infection. Although either of these mechanisms could render the APC inactive, it was unclear whether these processes are linked or represent independent (or redundant) pathways. The causative factor(s) in mediating these events and the question of whether such a factor(s) was of cellular or viral origin also remained unresolved.On the basis of the results of several recent studies (26, 32, 62), the viral protein kinase UL97 emerged as a likely candidate for involvement in the phosphorylation of Cdh1. Conserved among herpesviruses, UL97 functions in viral genome replication (7, 32, 81) and in nuclear egress of viral capsids (21, 39, 48). UL97 is present in the tegument of the virus particle (76) and is also expressed de novo with early kinetics (i.e., detectable by 5 h p.i. by Western blot assay), with increased expression at later times of the infection (51, 76, 77). UL97 is a serine/threonine (S/T) protein kinase (22), and recent studies have further characterized it as a Cdkl mimic, with predicted structural similarity to Cdk2 (64) and common substrates. UL97 has been shown to phosphorylate in vitro nuclear lamin A/C (21), the carboxyl-terminal domain of RNA polymerase II (5), the translation elongation factor 1δ (EF1δ) (33), and Rb (26, 62) on sites targeted by Cdks, and there is considerable evidence that UL97 phosphorylates lamin A/C, EF1δ, and Rb on these sites in infected cells as well (21, 26, 33, 62). Given that cyclin A/Cdk2 and cyclin B/Cdk1 complexes normally phosphorylate Cdh1, thus preventing its association with the APC, we hypothesized that UL97 phosphorylates Cdh1 during HCMV infection.In the present study, we provide further mechanistic details of the events and players involved in inactivating the APC during HCMV infection. Evidence that UL97 is the viral factor mediating the phosphorylation of Cdh1 was obtained. However, APC disassembly still occurred at similar times in ΔUL97 and wild-type virus infections, indicating that UL97-mediated phosphorylation of Cdh1 is not required for this event. The inactivation of the APC core complex is further attributed to the loss of the APC5 and APC4 subunits early during the infection. The degradation of these subunits is proteasome dependent and requires de novo synthesis of viral early or cellular proteins. While the primary mechanism of inactivation appears to be the dissociation of the complex and the targeted loss of APC5 and APC4, phosphorylation of Cdh1 may provide a small kinetic advantage and backup mechanism for disabling the APC.  相似文献   

18.
Herpes simplex virus (HSV) entry into cells is triggered by the binding of envelope glycoprotein D (gD) to a specific receptor, such as nectin-1 or herpesvirus entry mediator (HVEM), resulting in activation of the fusion effectors gB and gH and virus penetration. Here we report the identification of a hyperactive gB allele, D285N/A549T, selected by repeat passage of a gD mutant virus defective for nectin-1 binding through cells that express a gD-binding-impaired mutant nectin-1. The gB allele in a wild-type virus background enabled the use of other nectins as virus entry receptors. In addition, combination of the mutant allele with an epidermal growth factor receptor (EGFR)-retargeted gD gene yielded dramatically increased EGFR-specific virus entry compared to retargeted virus carrying wild-type gB. Entry of the gB mutant virus into nectin-1-bearing cells was markedly accelerated compared to that of wild-type virus, suggesting that the gB mutations affect a rate-limiting step in entry. Our observations indicate that ineffective gD activation can be complemented by hypersensitization of a downstream component of the entry cascade to gD signaling.Entry of herpes simplex virus type 1 (HSV-1) into susceptible cells involves the coordinated activities of at least five viral envelope glycoproteins (9, 18, 33, 40). Virions initially bind to glycosaminoglycan (GAG) moieties of cell surface proteoglycans through glycoproteins B and C (gB and gC, respectively) (32, 51), facilitating the interaction of gD with one of its specific receptors, herpesvirus entry mediator (HVEM, or HveA), nectin-1 (HveC), or 3-O-sulfated heparan sulfate (24, 45, 50). Receptor binding is believed to result in a conformational change in gD, which in turn activates the fusion mechanism mediated by gB and the gH/gL heterodimer; fusion merges the virus envelope with the cell surface or endosomal membrane, resulting in capsid release into the cytoplasm (11, 23, 30, 37, 44, 47, 48). Prior to receptor binding, the N-terminal region of the gD ectodomain is folded back over the immunoglobulin (Ig)-like core domain in a position to engage the C-terminal effector region (pro-fusion domain), thereby keeping the effector domain in an inactive state (23, 37). Receptor binding disrupts this engagement and liberates the effector domain to activate gB and/or gH/gL. The crystal structure of the gB ectodomain shows unexpected homology to the postfusion form of glycoprotein G from vesicular stomatitis virus (VSV G), a well-characterized fusion protein (30), providing strong evidence that gB plays a major role in membrane fusion. In addition, gH displays structural hallmarks of fusion proteins (26, 27), and gB and gH each have fusogenic activity, as indicated by the finding that either alone is sufficient for membrane fusion during nuclear egress (20). However, gB and gH/gL are both required for complete fusion during virus entry, although gB is dispensable for hemifusion, an intermediate state (53).Results from biochemical and bimolecular-complementation assays have shown that gD binds individually to both gB and gH/gL, regardless of the presence of gD receptors (4, 5, 25), while complexes of gB and gH/gL assemble only in the presence of receptor-bound gD (4, 5). These observations suggested that receptor-dependent gD activation brings gB and gH/gL together for execution of the fusion event. However, based on new evidence that gB and gH/gL can also interact in the absence of gD, an alternative model has been proposed in which activated gD signals to preformed gB-gH/gL complexes (6). While these models are not mutually exclusive, the functional significance of the detected complexes remains to be firmly established (15). However, there is broad consensus that the gD-receptor interaction triggers the initiation of fusion by direct interaction with either or both gB and gH/gL, indicating that the quality of the gD-receptor interaction is key to the efficiency of HSV infection.Viruses have an intrinsic ability to evolve and adapt to changes in the environment, including the acquisition of an extended host range which can lead to epidemic infections (56). We previously described gain-of-function derivatives of a gD mutant virus, K26-gD:R222N/F223I, that was impaired in its ability to use nectin-1 as an entry receptor (54). Repeated passage of this virus through cells that express nectin-1 as the sole gD receptor yielded phenotypic revertants that had regained the ability to use nectin-1 for infection. This phenotype resulted from reversion or forward mutations at the parental mutant positions or from substitutions elsewhere in gD that likely affect the integrity of the discontinuous interface with nectin-1. Since these types of experiments can reveal novel factors or interactions that are important for virus entry, we performed a similar study at higher stringency in an attempt to avoid simple reversion mutations. The strategy was to use our previous gD:R222N/F223I mutant virus that is defective for entry via nectin-1 and ask if this virus could adapt to host cells expressing a mutant form of nectin-1 whose binding to wild-type gD is severely impaired. A specific goal of this effort was to find mutations in gD or other envelope glycoproteins that could enhance infection through atypical receptors, including cell-type-specific receptors that can be engaged by retargeted HSV vectors.Here we report the identification of a hyperactive gB double mutation, gB:D285N/A549T, referred to herein as gB:N/T, that allows virus entry in the absence of authentic gD receptors, enhances virus entry through unconventional receptors, including a targeted receptor, and appears to act by sensitizing gB to activation by gD, directly or indirectly via gH/gL, and increasing the rate of virus entry into different host cells. Our observations demonstrate that hyperactive gB can compensate for ineffective gD-receptor interactions in the process of HSV entry into cells.  相似文献   

19.
20.
Hantaviruses infect endothelial cells and cause 2 vascular permeability-based diseases. Pathogenic hantaviruses enhance the permeability of endothelial cells in response to vascular endothelial growth factor (VEGF). However, the mechanism by which hantaviruses hyperpermeabilize endothelial cells has not been defined. The paracellular permeability of endothelial cells is uniquely determined by the homophilic assembly of vascular endothelial cadherin (VE-cadherin) within adherens junctions, which is regulated by VEGF receptor-2 (VEGFR2) responses. Here, we investigated VEGFR2 phosphorylation and the internalization of VE-cadherin within endothelial cells infected by pathogenic Andes virus (ANDV) and Hantaan virus (HTNV) and nonpathogenic Tula virus (TULV) hantaviruses. We found that VEGF addition to ANDV- and HTNV-infected endothelial cells results in the hyperphosphorylation of VEGFR2, while TULV infection failed to increase VEGFR2 phosphorylation. Concomitant with the VEGFR2 hyperphosphorylation, VE-cadherin was internalized to intracellular vesicles within ANDV- or HTNV-, but not TULV-, infected endothelial cells. Addition of angiopoietin-1 (Ang-1) or sphingosine-1-phosphate (S1P) to ANDV- or HTNV-infected cells blocked VE-cadherin internalization in response to VEGF. These findings are consistent with the ability of Ang-1 and S1P to inhibit hantavirus-induced endothelial cell permeability. Our results suggest that pathogenic hantaviruses disrupt fluid barrier properties of endothelial cell adherens junctions by enhancing VEGFR2-VE-cadherin pathway responses which increase paracellular permeability. These results provide a pathway-specific mechanism for the enhanced permeability of hantavirus-infected endothelial cells and suggest that stabilizing VE-cadherin within adherens junctions is a primary target for regulating endothelial cell permeability during pathogenic hantavirus infection.Hantaviruses cause 2 human diseases: hemorrhagic fever with renal syndrome (HFRS) and hantavirus pulmonary syndrome (HPS) (50). HPS and HFRS are multifactorial in nature and cause thrombocytopenia, immune and endothelial cell responses, and hypoxia, which contribute to disease (7, 11, 31, 42, 62). Although these syndromes sound quite different, they share common components which involve the ability of hantaviruses to infect endothelial cells and induce capillary permeability. Edema, which results from capillary leakage of fluid into tissues and organs, is a common finding in both HPS and HFRS patients (4, 7, 11, 31, 42, 62). In fact, both diseases can present with renal or pulmonary sequelae, and the renal or pulmonary focus of hantavirus diseases is likely to result from hantavirus infection of endothelial cells within vast glomerular and pulmonary capillary beds (4, 7, 11, 31, 42, 62). All hantaviruses predominantly infect endothelial cells which line capillaries (31, 42, 44, 61, 62), and endothelial cells have a primary role in maintaining fluid barrier functions of the vasculature (1, 12, 55). Although hantaviruses do not lyse endothelial cells (44, 61), this primary cellular target underlies hantavirus-induced changes in capillary integrity. As a result, understanding altered endothelial cell responses following hantavirus infection is fundamental to defining the mechanism of permeability induced by pathogenic hantaviruses (1, 12, 55).Pathogenic, but not nonpathogenic, hantaviruses use β3 integrins on the surface of endothelial cells and platelets for attachment (19, 21, 23, 39, 46), and β3 integrins play prominent roles in regulating vascular integrity (3, 6, 8, 24, 48). Pathogenic hantaviruses bind to basal, inactive conformations of β3 integrins (35, 46, 53) and days after infection inhibit β3 integrin-directed endothelial cell migration (20, 46). This may be the result of cell-associated virus (19, 20, 22) which keeps β3 in an inactive state but could also occur through additional regulatory processes that have yet to be defined. Interestingly, the nonpathogenic hantaviruses Prospect Hill virus (PHV) and Tula virus (TULV) fail to alter β3 integrin functions, and their entry is consistent with the use of discrete α5β1 integrins (21, 23, 36).On endothelial cells, αvβ3 integrins normally regulate permeabilizing effects of vascular endothelial growth factor receptor-2 (VEGFR2) (3, 24, 48, 51). VEGF was initially identified as an edema-causing vascular permeability factor (VPF) that is 50,000 times more potent than histamine in directing fluid across capillaries (12, 14). VEGF is responsible for disassembling adherens junctions between endothelial cells to permit cellular movement, wound repair, and angiogenesis (8, 10, 12, 13, 17, 26, 57). Extracellular domains of β3 integrins and VEGFR2 reportedly form a coprecipitable complex (3), and knocking out β3 causes capillary permeability that is augmented by VEGF addition (24, 47, 48). Pathogenic hantaviruses inhibit β3 integrin functions days after infection and similarly enhance the permeability of endothelial cells in response to VEGF (22).Adherens junctions form the primary fluid barrier of endothelial cells, and VEGFR2 responses control adherens junction disassembly (10, 17, 34, 57, 63). Vascular endothelial cadherin (VE-cadherin) is an endothelial cell-specific adherens junction protein and the primary determinant of paracellular permeability within the vascular endothelium (30, 33, 34). Activation of VEGFR2, another endothelial cell-specific protein, triggers signaling responses resulting in VE-cadherin disassembly and endocytosis, which increases the permeability of endothelial cell junctions (10, 12, 17, 34). VEGF is induced by hypoxic conditions and released by endothelial cells, platelets, and immune cells (2, 15, 38, 52). VEGF acts locally on endothelial cells through the autocrine or paracrine activation of VEGFR2, and the disassembly of endothelial cell adherens junctions increases the availability of nutrients to tissues and facilitates leukocyte trafficking and diapedesis (10, 12, 17, 55). The importance of endothelial cell barrier integrity is often in conflict with requirements for endothelial cells to move in order to permit angiogenesis and repair or cell and fluid egress, and as a result, VEGF-induced VE-cadherin responses are tightly controlled (10, 17, 18, 32, 33, 59). This limits capillary permeability while dynamically responding to a variety of endothelial cell-specific factors and conditions. However, if unregulated, this process can result in localized capillary permeability and edema (2, 9, 10, 12, 14, 17, 29, 60).Interestingly, tissue edema and hypoxia are common findings in both HPS and HFRS patients (11, 31, 62), and the ability of pathogenic hantaviruses to infect human endothelial cells provides a means for hantaviruses to directly alter normal VEGF-VE-cadherin regulation. In fact, the permeability of endothelial cells infected by pathogenic Andes virus (ANDV) or Hantaan virus (HTNV) is dramatically enhanced in response to VEGF addition (22). This response is absent from endothelial cells comparably infected with the nonpathogenic TULV and suggests that enhanced VEGF-induced endothelial cell permeability is a common underlying response of both HPS- and HFRS-causing hantaviruses (22). In these studies, we comparatively investigate responses of human endothelial cells infected with pathogenic ANDV and HTNV, as well as nonpathogenic TULV.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

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