Cytoskeleton during Late Steps of HIV-1 Life Cycle: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor: , , , , , ,

HIV-1 has evolved a plethora of strategies to overcome the cytoskeletal barrier (i.e., actin and intermediate filaments (AFs and IFs) and microtubules (MTs)) to achieve the viral cycle. HIV-1 modifies cytoskeletal organization and dynamics by acting on associated adaptors and molecular motors to productively fuse, enter, and infect cells and then traffic to the cell surface, where virions assemble and are released to spread infection. The HIV-1 envelope (Env) initiates the cycle by binding to and signaling through its main cell surface receptors (CD4/CCR5/CXCR4) to shape the cytoskeleton for fusion pore formation, which permits viral core entry. Then, the HIV-1 capsid is transported to the nucleus associated with cytoskeleton tracks under the control of specific adaptors/molecular motors, as well as HIV-1 accessory proteins. Furthermore, HIV-1 drives the late stages of the viral cycle by regulating cytoskeleton dynamics to assure viral Pr55Gag expression and transport to the cell surface, where it assembles and buds to mature infectious virions. 

  • HIV-1 infection
  • HIV-1 Env complex
  • HIV-1 Pr55Gag
  • cytoskeleton dynamics
  • cytoskeleton adaptors and motor proteins

1. Actin Filaments

In HIV-1 infection, cell-to-cell viral transmission at virological synapse (VS) has been well documented and is mainly responsible for the high decrease in T CD4+ lymphocytes during the acute phase but especially during the AIDS phase of infection, thus having clinical importance [1]. This manner of infection involves different pathways but includes events in which an infected cell makes contact with a noninfected cell, transmitting the virus in different ways: during VS (transmission or transfer) and by cytoskeletal structures, such as filopodia or nanotubes [1][2][3]. Hence, during viral egress, AFs stand out as scaffolds for the virus to bud. The plasma membrane is able to form structures such as filopodia or nanotubes for cell-to-cell viral transmission, as filopodia extend through receptor-mediated mechanisms from uninfected cells towards the infected cell, and a similar retroviral surfing has been described over nanotubes that also transfer viral proteins inside [2][4][5][6][7][8][9][10][11][12]. In this case, receptor specificity seems to play a minor role, as these actin-driven structures seem to extend from HIV-1-infected cells to target cells in the absence of receptor–Env interactions [2][4][7][8][9][10][11][13]. In DCs, by budding at the tip of filopodia, HIV-1 can tether several neighboring CD4+ T cells, leading to viral transfer and infection of the targeted T cells [14]. Nanopores are structures formed between cells, approximately 100 nm in length, and depend on microfilament structures [15] that, in the context of HIV-1 VS, are formed in the membrane of the target cell in which the virus “surfs” through that structure to infect and in an irrespective manner of receptor–HIV-1 Env interaction [5]. Infection-mediated nanotubes could be established between T CD4+ lymphocytes (infected and noninfected pairs) or between T CD4+ lymphocytes and dendritic cells (DCs) [5]. In this matter, during the infection of DCs, HIV-1 (which enters the cell using the lectin DC-SIGN) mediates the activation of a GTPase and the remodeling of the actin cytoskeleton to promote filopodia extension that allows virus transmission to neighboring CD4+ T cells [5][14].
The ERM actin-adaptors also have relevance in HIV-1 production [16], since these proteins connect cortical F-actin with integral and peripheral membrane proteins they are incorporated into virions, interacting with cellular components of the virological presynapse. Phosphorylation activates ezrin, which specifically accumulates at the HIV-1 presynapse in T-cell lines and primary CD4+ lymphocytes [16]. Moreover, while cells did not tolerate a complete knockdown of ezrin, even a modest reduction in ezrin expression of approximately 50% in HIV-1-producing cells led to the release of particles with impaired infectivity. Furthermore, when cocultured with uninfected target cells, ezrin-knockdown producer cells displayed reduced accumulation of the tetraspanin CD81 at the VS and formed syncytia. Such an outcome is likely not optimal for virus dissemination, as evidenced by the fact that, in vivo, only relatively few infected cells form syncytia. Thus, ezrin likely helps secure efficient virus spread not only by enhancing virion infectivity but also by preventing excessive membrane fusion at the VS [16]. Likewise, moesin regulates HIV-1 Env/CD4-mediated pore fusion formation in a cell-to-cell VS model, where moesin phosphorylation and its actin/plasma membrane-anchoring activity are required [17].
Recently, the factor EWI-2, a protein that was previously shown to associate with ezrin and tetraspanins, has been identified as a host factor that contributes to the inhibition of HIV-1 Env-mediated cell–cell fusion [18]. It has been observed that EWI-2 accumulates at the presynaptic terminal where it contributes to the fusion-preventing activities of the other viral and cellular components, and EWI-2 was also downregulated upon HIV-1 infection, most likely by the accessory viral protein U (Vpu) [18]. In this sense, the expression levels of EWI-2 and CD81 molecules are restored on the surface of HIV-1 Env-driven syncytia [18], where they act as fusion inhibitors, thereby providing novel insight into how deathly syncytia (i.e., collapsing by apoptosis [19][20][21]) might be prevented from fusing indefinitely. This ‘syncytial apoptosis’ is frequently detected during HIV-1 infection in vitro and in vivo in tissues from HIV-1-infected patients [20]. Hence, and in contrast to potential EWI-2/CD81 inhibitors [18], it has been proposed that the induction of selective death of HIV-1-elicited syncytia might lead to the elimination of viral reservoirs. Thus, specific drugs or strategies aligned with this concept would constitute a therapeutic complement to current ARTs [21].
One key factor that could associate cortical AFs with the cell surface and appears to regulate late stages of the HIV-1 infection cycle is the interferon (IFN)-induced type II membrane glycoprotein tetherin, also known as bone marrow stromal cell antigen 2 (BST2), HM1.24 or CD317 [22][23]. Tetherin is a GPI anchor protein localized in lipid raft microdomains that can link rafts with the underlying cortical actin cytoskeleton [24] and acts as a direct physical tether for virions, linking the HIV-1 Env complex to the plasma membrane and preventing the final “escape” of viruses from the infected producer cell [25][26][27][28][29][30]. Tetherin thus severely limits the spread of cell-free viruses and reduces the cell-to-cell spread of HIV-1 to a lesser degree [31][32]. HIV-1 uses its Vpu protein to counteract the BST2/tetherin barrier [33][34]. Vpu localizes predominantly to the trans-Golgi network (TGN) and recycling endosomes [35][36]. Thus, Vpu antagonizes tetherin in a post-ER compartment [37], downregulating BST2/tetherin from the cell-surface [34] by recruiting a βTRCP2-SCF-Cullin1 ubiquitin ligase complex [38][39][40][41][42][43].
P-selectin glycoprotein ligand 1 (PSGL-1) is another important factor recently related to HIV-1 replication that could restrict actin dynamics and arrest viral Env at the plasma membrane, resulting in virions with poor Env incorporation and reduced infectivity [44]. In fact, PSGL-1 is a dimeric mucin-like glycoprotein that binds to P-, E-, and L-selectins and plays a role in leukocyte rolling on endothelial surfaces prior to transmigration [45][46][47][48][49][50], being identified as an IFN-γ-induced restriction factor in CD4+ T cells using a genome-wide proteomic screen [51]. In the context of HIV-1 infection, PSGL-1 seems to be recruited to the sites where HIV-1 particles assemble [52]. Likewise, it seems that PSGL-1 inhibits reverse transcription in target cells and diminishes the infectivity of released virions. This inhibitory effect could be antagonized by the HIV-1 Vpu protein through the ubiquitination and proteasomal degradation of PSGL-1. Therefore, PSGL-1 could be incorporated into HIV-1 virions, exerting a negative effect on HIV-1 transmission, not only by inhibiting Env processing and incorporation but also by directly inhibiting virion attachment to target cells [53].
A recent study concluded that HIV-1 molds AFs cortical nodes in areas of high positive membrane curvature within Arp2/3-dependent F-actin filopodia and lamellipodia structures, which enables HIV-1 to bud at the cell edge [54]. This research is supported by live cell imaging and focused ion beam scanning electron microscopy (FIB-SEM), which permit the observation of F-actin structures that exhibit strong positive curvature where HIV-1 buds [54]. Likewise, virion proteomics, gene silencing, and viral mutagenesis supported not only the functional involvement of Arp2/3 but also the involvement of the Rac1- and Cdc42-IQGAP1 pathways in driving F-actin regulation and membrane curvature, thus allowing HIV-1 to egress. HIV-1 also activates the Cdc42-Arp2/3 filopodial pathway for release, requiring the scaffolding protein IQGAP1 (a Rac1 and Cdc42 binding partner) [55][56], a process that similarly occurs during cell-to-cell viral spreading [54].
The HIV-1 structural Pr55Gag polyprotein could be coimmunoprecipitated with actin but not tubulin, where a direct physical interaction between HIV-1 Pr55Gag and F-actin via the nucleocapsid (NC) domain of Pr55Gag has been reported [57]. Likewise, it is known that there is a physical and tight association between HIV-1 Pr55Gag and actin both in fixed and adherent cell lines [58]. In addition, several studies have observed a high association of F-actin with HIV-1 budding sites, as some researchers have observed NC-dependent actin structures that emanate from HIV-1 buds during viral assembly and disappear upon viral release [59]. Additionally, the presence of AFs near budding sites was later confirmed and proven to have direct physical contact with nascent viral particles [60][61] (Figure 1). However, some observations argue against a functional role of these microfilaments in viral budding [61][62], but since these studies did not measure the actin content in cell-free viral supernatants, a role for actin in HIV-1 egress could not be ruled out. Therefore, given the importance of viral RNA binding to the NC region of the Pr55Gag polyprotein during the viral life cycle, functional characterization of the NC-actin association and how this association is influenced by the viral RNA+ genome should be further addressed to understand the role played by AFs in HIV-1 Pr55Gag packaging and virion release.
Figure 1. Scheme illustrating the late stages of the HIV-1 life cycle and associated tubulin and actin cytoskeleton dynamics. The transcription and translation of the integrated HIV-1 genome generate viral RNA+ and proteins, with the structural Pr55Gag polyprotein being relevant to recognize viral RNA+ and recruit a complex to stable MTs to travel to the plasma membrane (right panel). In fact, HIV-1 Pr55Gag uses MTs-dependent cellular machinery to traffic to the cell surface in a SOCS1-dependent manner. Pr55Gag also interacts with F-actin, thus assembling at the cell surface in areas where AFs are remodeled by the Arp2/3- and Rac1/Cdc42/IQGAP1 pathways, inducing F-actin filopodia and lamellipodia structures that finally generate high positive membrane curvatures where HIV-1 egresses (right panel). The MT-associated HDAC6 enzyme deacetylates stable MTs, thereby impeding Pr55Gag cell surface localization (left panel). Likewise, HDAC6 tubulin-deacetylase function is required to target HIV-1 Pr55Gag and Vif proteins for autophagy degradation, thus inhibiting viral production and infection capacity (left panel). HIV-1 Nef counteracts these antiviral actions by targeting the MT-associated enzyme, thereby stabilizing MTs and the HIV-1 Pr55Gag and Vif proteins. Therefore, HIV-1 Nef promotes the transport of Pr55Gag to the cell surface where it assembles to form viral particles that incorporate Vif to ensure the infectivity of HIV-1. Designs and templates were created with BioRender.
A summary of the main data discussed in this section is presented in Table 1.
Table 1. Cellular and viral factors that influence actin cytoskeletal dynamics during the late stages of the HIV-1 viral cycle.
Cell Factor Function of the Cellular Factor Impact on HIV-1 Infection 1
ERM Secure efficient virus spread not only by enhancing virion infectivity but also by preventing excessive membrane fusion at the virological synapse (VS). +
EWI-2/CD81 Difficult HIV-1 Env-mediates cell-to-cell fusion.
BST2/tetherin Links AFs with plasma membrane where it sequesters the viral Env and preventing the virus escape from cell.
PSGL-1 Inhibits RT in target cells and diminished the infectivity of released virions.
Arp2/3 and Rac-1/Cdc42- IQGAP1 pathways Remodeling of cortical AFs in areas of high positive membrane curvature, within Arp2/3- and Rac1/Cdc42/IQGAP1-dependent F-actin filopodia and lamellipodia structures, which enables HIV-1 to bud and cell-to-cell spreading. +
AFs Cortical AFs bind HIV-1 Pr55Gag protein by the NC domain. +
Viral factor Function of the viral factor Impact on HIV-1 infection 1
Pr55Gag Actin interaction facilitating viral egress. +
Vpu Counteracts BST-2/tetherin, EWI-2, and PSGL-1 antiviral activities. +
1 (+) represents the promotion of HIV-1 replication. (−) represents inhibition of HIV-1 replication.

2. Microtubules

It is thought that HIV-1 release and cell-to-cell transfer rely on intact and stable acetylated MTs [63][64][65][66], such as through VS, where during transfer, the MTOC in HIV-1-infected T cells, but not in target cells, polarizes towards the VS [65][66][67]. Notably, in infected peripheral blood lymphocytes (PBLs), the stabilization of MTs by chemical inhibition of the tubulin-deacetylase HDAC6 increases HIV-1 viral production and replication for several days postinfection [63]. This fact indicates the importance of acetylated stable MTs for the late stages of the HIV-1 life cycle and points to the regulatory role of HDAC6 in HIV-1 replication.
In this sense, the tubulin-deacetylase HDAC6 appears to be crucial in the control of HIV-1 production and the infection capacity of nascent viral particles [63][68][69][70]. HDAC6 directly interacts with A3G (apolipoprotein B mRNA-editing enzymatic polypeptide-like 3G or APOBEC3G) to undergo cellular codistribution along MTs and the cytoplasm in an A3G/HDAC6 complex [70]. HDAC6 competes for Vif-mediated A3G degradation, also accounting for the steady-state A3G expression level. In fact, HDAC6 directly interacts with and promotes Vif autophagic clearance due to its C-terminal BUZ domain, a process requiring the tubulin-deacetylase activity of HDAC6 [70]. HDAC6 degrades Vif but does not affect core-binding factor β (CBF-β), a Vif-associated partner reported to be key for Vif-mediated A3G degradation. Thus, HDAC6 antagonizes the proviral activity of the Vif/CBF-β-associated complex by targeting Vif and stabilizing A3G. Finally, in cells producing virions, a good correlation was observed between the ability of HDAC6 to degrade Vif and restore A3G expression, suggesting that HDAC6 controls the amount of Vif incorporated into nascent virions and the ability of HIV-1 particles to be infectious [70]. The tubulin-deacetylase HDAC6 also promotes the autophagic degradation of the viral polyprotein Pr55Gag to inhibit HIV-1 production [69]. The HIV-1 Nef factor counteracts this antiviral activity of HDAC6 by inducing its cellular clearance and subsequently stabilizing Pr55Gag and Vif viral proteins. HIV-1 Nef interacts with HDAC6, colocalizing on MTs and neutralizing HDAC6 by degradation in an acidic/endosomal-lysosomal pathway [69]. HIV-1 Nef therefore assures Pr55Gag location and aggregation at the plasma membrane, promoting virus production and enhancing the infectivity of viral particles by the stabilization and uptake of Vif [69]. Moreover, a recent study showed that the overexpression of TDP-43 in virus-producing cells stabilizes HDAC6 (at the mRNA and protein levels) and triggers the autophagic clearance of HIV-1 Pr55Gag and Vif proteins [68]. These events inhibit viral particle production and impair virion infectiveness, resulting in a reduction in the amount of Pr55Gag and Vif proteins incorporated into virions. [68]. A nuclear localization signal (NLS)-TDP-43 mutant is not able to control HIV-1 viral production and infection. Likewise, specific TDP-43 knockdown reduces HDAC6 expression (i.e., mRNA and protein) and increases the expression levels of HIV-1 Vif and Pr55Gag proteins, thereby stabilizing MTs by α-tubulin acetylation [68]. Thus, TDP-43 silencing favors virion production and enhances virus infectious capacity, thereby increasing the amount of Vif and Pr55Gag proteins incorporated into virions. Notably, there is a direct relationship between the content of Vif and Pr55Gag proteins in virions and their infection capacity [68] (Figure 1).
Furthermore, the concomitant PTM of MTs by acetylation produced after silencing the HDAC6/TDP-43 axis could stabilize Pr55Gag, facilitating its transport to the plasma membrane. In this sense, it has been reported that the Pr55Gag polyprotein colocalizes, early after expression, with viral RNA through the cis-acting packaging element, psi (Ψ), in the perinuclear region and centrioles. Ψ+ RNA and Pr55Gag are subsequently transported to the plasma membrane of the cell [58]. Moreover, HIV-1 Pr55Gag uses MT-dependent cellular machinery to traffic to the cell surface in a manner dependent on the suppressor of cytokine signaling 1 (SOCS1) [71], where stable MTs are key for SOCS1/Pr55Gag transport [71]. SOCS1 is a negative regulator of cytokine signaling [72] and interacts with the MA and NC regions of Pr55Gag through its central SH2 domain to facilitate Pr55Gag intracellular trafficking and stability, thereby regulating the late stages of the virus replication pathway [73] (Figure 1).
In contrast, some MTs-associated molecules, such as the IQGAP family of scaffold proteins, which also regulate the actin cytoskeleton [21][74][75][76], have been reported to exert an inhibitory effect on HIV-1 budding and maturation by targeting Pr55Gag [77]. IQGAP1 has been shown to interact with the budding of enveloped viruses, as is the case for HIV-1 [78][79][80]. Overexpression of IQGAP1 inhibits HIV-1 budding, while depletion of IQGAP1 enhances HIV-1 particle release [77]. In fact, IQGAP1 directly interacts with both the NC and p6 regions of the structural Pr55Gag viral protein, altering and impairing its distribution at the plasma membrane, thereby being responsible for the inhibition of viral release [77] (Figure 1).
An example of HIV-1 viral factors that play an important role in the late stages of the viral cycle and interact with tubulin cytoskeleton are the accessory Tat (transactivator of transcription) and viral protein R (Vpr) proteins [81][82].
It has been observed that HIV-1 Tat interacts with monomeric tubulin and MTs. This Tat–MT association does not appear to alter or be required for Tat secretion or uptake. However, the association of Tat with MTs induces apoptosis, as the combination of Tat and MT leads to the alteration of MT dynamics and activation of a mitochondria-dependent apoptotic pathway [81]. In addition, Bim facilitates Tat-induced apoptosis, since Bim is a proapoptotic Bcl-2 relative and a transducer of death signals initiated by perturbation of MT dynamics [81]. Therefore, the proapoptotic Tat–MT interplay could account, in part, for the immune cell dysfunctions and pathogenesis associated with HIV-1 infection. HIV-1 Vpr is able to interact with MT-associated factors, such as EB1, p150Glued, and dynein heavy chain [82], and alters the MT plus end localization of EB1 and p150Glued, negatively affecting the centripetal movement of phagosomes and their maturation [82]. Moreover, Vpr appears to target MT/dynein-dependent endocytic trafficking in HIV-1-infected macrophages, which deeply alters phagolysosome biogenesis [82]. It is plausible that this Vpr-provoked cell dysfunction could be toxic for infected cells by acting on caspases [83][84][85], and mitochondria-associated factors (reviewed in [86][87][88]) that could be activated after phagolysosome disruption [89].
A summary of the main data discussed in this section is presented in Table 2.

3. Intermediate Filaments

The distribution of vimentin IFs around the nucleus and their extension throughout the cytoplasm provide a scaffold for cellular components (reviewed in [90]) and allow vimentin to play an essential role in cellular cargo transportation [91][92][93][94][95][96][97][98][99][100], as well as in the trafficking of viral components to the location of cell assembly and egress (reviewed in [101][102]). It is interesting to note that vimentin IFs are distributed in cells by MTs-dependent dynamic transport [103][104][105][106][107], regulated by AFs and related kinases [108], which, in turn, is required to maintain the cytoskeleton network [109][110]. Vimentin IFs modulate the replication, assembly, and egress of viruses in the host due to their known function of regulating endosomal trafficking via Rab7a and Polo-like kinase 1 (Plk1). Rab7a, which is ubiquitously present in early and late endosomes [111][112], interacts with insoluble and soluble vimentin [113][114]. Rab7a interacts directly with vimentin, and this interaction modulates vimentin phosphorylation and assembly [113]. Rab7a-depleted cells have an abundance of insoluble vimentin IFs and defective endosomal trafficking [115]. Phosphorylation of vimentin IFs at Ser459 by Polo-like kinase 1 (Plk1) inhibits endolytic fusion during mitosis [96]. Altogether, these interactions demonstrate that vimentin is a critical regulator of late endocytic trafficking and egress of viral particles [116][117][118][119].
In the context of HIV-1, the knockdown of the vimentin protein has been associated with a reduction in the HIV-1 p24 capsid (CA) protein (derived from Pr55Gag) level with concomitant impairment of viral replication [120]. Likewise, the intracellular administration of a peptide that modifies vimentin IF distribution similarly inhibits HIV-1 replication [120]. Moreover, the interplay of vimentin IFs with the Mac-2-binding protein (M2BP [121]), a protein that is secreted in HIV-1-infected patients and correlates with viremia progression or control [122][123][124], mediates the association of vimentin IFs with HIV-1 Pr55Gag [125]. Thus, vimentin/M2BP interplay negatively affects the trafficking of HIV-1 Pr55Gag towards the plasma membrane, whereas the absence of M2BP favors the cell surface localization of the HIV-1 Pr55Gag polyprotein [125][126].
It is interesting to note that some clinical studies with HIV-1 patients suggest vimentin as a potential marker for ART treatment efficiency, since in HIV-1-positive peripheral blood mononuclear cells (PBMCs) isolated from patients, vimentin was verified to be related to ART treatment and viremia control [127]. Although there are certain limitations of this research, the reported data could be of interest. For example, the protein–protein interaction analysis was database dependent, and limited clinical samples were obtained to verify the expression level of vimentin in plasma under different experimental conditions [127]. Moreover, several reports point out that some viruses could rearrange vimentin filaments (reviewed in [101]), and in the case of HIV-1, some published data support the possibility that vimentin may serve as a substrate for HIV-1 proteins within infected cells [128]. In this sense, the IFs vimentin, desmin, and glial fibrillary acidic proteins are cleaved in vitro by the HIV-1 PR enzyme [128]. Thus, microsequencing analysis showed that HIV-1 PR cleaved both human and murine vimentin. Microinjection of HIV-1 PR into cultured human fibroblasts resulted in a 9-fold increase in the percentage of cells with an altered and abnormal distribution of vimentin IFs [128]. The ability of the HIV-1 PR to liberate amino-terminal peptides from vimentin, representing one of the best substrates of the viral PR, may be important to achieve the replication cycle because PR-derived vimentin fragments disrupt IFs organization [129][130]. In the case of HIV-1 Vif, this accessory protein is found in the cytoplasm and in the perinuclear area colocalizing with vimentin, thereby disrupting the vimentin network [131]. In contrast, Vif presents a diffuse pattern of distribution in the cytoplasm, nuclear membrane and nucleus when the vimentin network is chemically blocked [131].
Most commonly, when the vimentin network is perturbed, their IFs collapse into a clump with a juxtanuclear localization. Similarly, amino-terminal polypeptides of vimentin are responsible for changes in nuclear architecture associated with HIV-1 PR activity in tissue culture cells isolated from infected individuals, and condensation or degeneration of nuclear chromatin has been described in cells from a variety of tissues [132]. Furthermore, infection with HIV-1 includes an increase in the incidence of cancer, particularly lymphomas. Attempts at identifying an additional agent responsible for non-Hodgkin’s lymphoma in a large HIV-1-infected cohort have failed [133], raising the possibility that HIV-1 itself is directly responsible. The ability of the HIV-1 PR to liberate amino-terminal peptides from vimentin may interfere with host cell gene expression [128][134] and may be important to achieve the replication cycle because PR-derived vimentin fragments disrupt IF organization [129][130]. In the case of HIV-1 Vif, this accessory protein is found in the perinuclear area colocalizing with vimentin, thereby disrupting the vimentin network [131]. Altogether, these data indicate that the relevance of the HIV-1 Vif/vimentin IF interplay for HIV-1 replication remains to be elucidated.

This entry is adapted from the peer-reviewed paper 10.3390/ijms241713104

References

  1. Puigdomenech, I.; Massanella, M.; Cabrera, C.; Clotet, B.; Blanco, J. On the steps of cell-to-cell HIV transmission between CD4 T cells. Retrovirology 2009, 6, 89.
  2. De Armas-Rillo, L.; Valera, M.S.; Marrero-Hernandez, S.; Valenzuela-Fernandez, A. Membrane dynamics associated with viral infection. Rev. Med. Virol. 2016, 26, 146–160.
  3. Bracq, L.; Xie, M.; Benichou, S.; Bouchet, J. Mechanisms for Cell-to-Cell Transmission of HIV-1. Front. Immunol. 2018, 9, 260.
  4. Barroso-Gonzalez, J.; Garcia-Exposito, L.; Puigdomenech, I.; de Armas-Rillo, L.; Machado, J.D.; Blanco, J.; Valenzuela-Fernandez, A. Viral infection: Moving through complex and dynamic cell-membrane structures. Commun. Integr. Biol. 2011, 4, 398–408.
  5. Sowinski, S.; Jolly, C.; Berninghausen, O.; Purbhoo, M.A.; Chauveau, A.; Kohler, K.; Oddos, S.; Eissmann, P.; Brodsky, F.M.; Hopkins, C.; et al. Membrane nanotubes physically connect T cells over long distances presenting a novel route for HIV-1 transmission. Nat. Cell Biol. 2008, 10, 211–219.
  6. Sherer, N.M.; Lehmann, M.J.; Jimenez-Soto, L.F.; Horensavitz, C.; Pypaert, M.; Mothes, W. Retroviruses can establish filopodial bridges for efficient cell-to-cell transmission. Nat. Cell Biol. 2007, 9, 310–315.
  7. Do, T.; Murphy, G.; Earl, L.A.; Del Prete, G.Q.; Grandinetti, G.; Li, G.H.; Estes, J.D.; Rao, P.; Trubey, C.M.; Thomas, J.; et al. Three-dimensional imaging of HIV-1 virological synapses reveals membrane architectures involved in virus transmission. J. Virol. 2014, 88, 10327–10339.
  8. Hashimoto, M.; Bhuyan, F.; Hiyoshi, M.; Noyori, O.; Nasser, H.; Miyazaki, M.; Saito, T.; Kondoh, Y.; Osada, H.; Kimura, S.; et al. Potential Role of the Formation of Tunneling Nanotubes in HIV-1 Spread in Macrophages. J. Immunol. 2016, 196, 1832–1841.
  9. Okafo, G.; Prevedel, L.; Eugenin, E. Tunneling nanotubes (TNT) mediate long-range gap junctional communication: Implications for HIV cell to cell spread. Sci. Rep. 2017, 7, 16660.
  10. Eugenin, E.A.; Gaskill, P.J.; Berman, J.W. Tunneling nanotubes (TNT): A potential mechanism for intercellular HIV trafficking. Commun. Integr. Biol. 2009, 2, 243–244.
  11. Eugenin, E.A.; Gaskill, P.J.; Berman, J.W. Tunneling nanotubes (TNT) are induced by HIV-infection of macrophages: A potential mechanism for intercellular HIV trafficking. Cell Immunol. 2009, 254, 142–148.
  12. Lehmann, M.J.; Sherer, N.M.; Marks, C.B.; Pypaert, M.; Mothes, W. Actin- and myosin-driven movement of viruses along filopodia precedes their entry into cells. J. Cell Biol. 2005, 170, 317–325.
  13. Sutton, J.; Dotson, D.; Dong, X. Molecular Events in Late Stages of HIV-1 Replication. JSM Microbiol. 2013, 1, 1004.
  14. Aggarwal, A.; Iemma, T.L.; Shih, I.; Newsome, T.P.; McAllery, S.; Cunningham, A.L.; Turville, S.G. Mobilization of HIV spread by diaphanous 2 dependent filopodia in infected dendritic cells. PLoS Pathog. 2012, 8, e1002762.
  15. Ljubojevic, N.; Henderson, J.M.; Zurzolo, C. The Ways of Actin: Why Tunneling Nanotubes Are Unique Cell Protrusions. Trends Cell Biol. 2021, 31, 130–142.
  16. Roy, N.H.; Lambele, M.; Chan, J.; Symeonides, M.; Thali, M. Ezrin is a component of the HIV-1 virological presynapse and contributes to the inhibition of cell-cell fusion. J. Virol. 2014, 88, 7645–7658.
  17. Barrero-Villar, M.; Cabrero, J.R.; Gordon-Alonso, M.; Barroso-Gonzalez, J.; Alvarez-Losada, S.; Munoz-Fernandez, M.A.; Sanchez-Madrid, F.; Valenzuela-Fernandez, A. Moesin is required for HIV-1-induced CD4-CXCR4 interaction, F-actin redistribution, membrane fusion and viral infection in lymphocytes. J. Cell Sci. 2009, 122, 103–113.
  18. Whitaker, E.E.; Matheson, N.J.; Perlee, S.; Munson, P.B.; Symeonides, M.; Thali, M. EWI-2 Inhibits Cell-Cell Fusion at the HIV-1 Virological Presynapse. Viruses 2019, 11, 1082.
  19. Andreau, K.; Perfettini, J.L.; Castedo, M.; Metivier, D.; Scott, V.; Pierron, G.; Kroemer, G. Contagious apoptosis facilitated by the HIV-1 envelope: Fusion-induced cell-to-cell transmission of a lethal signal. J. Cell Sci. 2004, 117, 5643–5653.
  20. Ferri, K.F.; Jacotot, E.; Geuskens, M.; Kroemer, G. Apoptosis and karyogamy in syncytia induced by the HIV-1-envelope glycoprotein complex. Cell Death Differ. 2000, 7, 1137–1139.
  21. Nardacci, R.; Perfettini, J.L.; Grieco, L.; Thieffry, D.; Kroemer, G.; Piacentini, M. Syncytial apoptosis signaling network induced by the HIV-1 envelope glycoprotein complex: An overview. Cell Death Dis. 2015, 6, e1846.
  22. Blasius, A.L.; Giurisato, E.; Cella, M.; Schreiber, R.D.; Shaw, A.S.; Colonna, M. Bone marrow stromal cell antigen 2 is a specific marker of type I IFN-producing cells in the naive mouse, but a promiscuous cell surface antigen following IFN stimulation. J. Immunol. 2006, 177, 3260–3265.
  23. Ishikawa, J.; Kaisho, T.; Tomizawa, H.; Lee, B.O.; Kobune, Y.; Inazawa, J.; Oritani, K.; Itoh, M.; Ochi, T.; Ishihara, K.; et al. Molecular cloning and chromosomal mapping of a bone marrow stromal cell surface gene, BST2, that may be involved in pre-B-cell growth. Genomics 1995, 26, 527–534.
  24. Rollason, R.; Korolchuk, V.; Hamilton, C.; Jepson, M.; Banting, G. A CD317/tetherin-RICH2 complex plays a critical role in the organization of the subapical actin cytoskeleton in polarized epithelial cells. J. Cell Biol. 2009, 184, 721–736.
  25. Fitzpatrick, K.; Skasko, M.; Deerinck, T.J.; Crum, J.; Ellisman, M.H.; Guatelli, J. Direct restriction of virus release and incorporation of the interferon-induced protein BST-2 into HIV-1 particles. PLoS Pathog. 2010, 6, e1000701.
  26. Perez-Caballero, D.; Zang, T.; Ebrahimi, A.; McNatt, M.W.; Gregory, D.A.; Johnson, M.C.; Bieniasz, P.D. Tetherin inhibits HIV-1 release by directly tethering virions to cells. Cell 2009, 139, 499–511.
  27. Hammonds, J.; Wang, J.J.; Yi, H.; Spearman, P. Immunoelectron microscopic evidence for Tetherin/BST2 as the physical bridge between HIV-1 virions and the plasma membrane. PLoS Pathog. 2010, 6, e1000749.
  28. Habermann, A.; Krijnse-Locker, J.; Oberwinkler, H.; Eckhardt, M.; Homann, S.; Andrew, A.; Strebel, K.; Krausslich, H.G. CD317/tetherin is enriched in the HIV-1 envelope and downregulated from the plasma membrane upon virus infection. J. Virol. 2010, 84, 4646–4658.
  29. Hammonds, J.; Wang, J.J.; Spearman, P. Restriction of Retroviral Replication by Tetherin/BST-2. Mol. Biol. Int. 2012, 2012, 424768.
  30. Venkatesh, S.; Bieniasz, P.D. Mechanism of HIV-1 virion entrapment by tetherin. PLoS Pathog. 2013, 9, e1003483.
  31. Giese, S.; Marsh, M. Tetherin can restrict cell-free and cell-cell transmission of HIV from primary macrophages to T cells. PLoS Pathog. 2014, 10, e1004189.
  32. Kuhl, B.D.; Sloan, R.D.; Donahue, D.A.; Bar-Magen, T.; Liang, C.; Wainberg, M.A. Tetherin restricts direct cell-to-cell infection of HIV-1. Retrovirology 2010, 7, 115.
  33. Neil, S.J.; Zang, T.; Bieniasz, P.D. Tetherin inhibits retrovirus release and is antagonized by HIV-1 Vpu. Nature 2008, 451, 425–430.
  34. Van Damme, N.; Goff, D.; Katsura, C.; Jorgenson, R.L.; Mitchell, R.; Johnson, M.C.; Stephens, E.B.; Guatelli, J. The interferon-induced protein BST-2 restricts HIV-1 release and is downregulated from the cell surface by the viral Vpu protein. Cell Host Microbe 2008, 3, 245–252.
  35. Dube, M.; Roy, B.B.; Guiot-Guillain, P.; Mercier, J.; Binette, J.; Leung, G.; Cohen, E.A. Suppression of Tetherin-restricting activity upon human immunodeficiency virus type 1 particle release correlates with localization of Vpu in the trans-Golgi network. J. Virol. 2009, 83, 4574–4590.
  36. Varthakavi, V.; Smith, R.M.; Martin, K.L.; Derdowski, A.; Lapierre, L.A.; Goldenring, J.R.; Spearman, P. The pericentriolar recycling endosome plays a key role in Vpu-mediated enhancement of HIV-1 particle release. Traffic 2006, 7, 298–307.
  37. Schubert, U.; Strebel, K. Differential activities of the human immunodeficiency virus type 1-encoded Vpu protein are regulated by phosphorylation and occur in different cellular compartments. J. Virol. 1994, 68, 2260–2271.
  38. Douglas, J.L.; Viswanathan, K.; McCarroll, M.N.; Gustin, J.K.; Fruh, K.; Moses, A.V. Vpu directs the degradation of the human immunodeficiency virus restriction factor BST-2/Tetherin via a betaTrCP-dependent mechanism. J. Virol. 2009, 83, 7931–7947.
  39. Goffinet, C.; Allespach, I.; Homann, S.; Tervo, H.M.; Habermann, A.; Rupp, D.; Oberbremer, L.; Kern, C.; Tibroni, N.; Welsch, S.; et al. HIV-1 antagonism of CD317 is species specific and involves Vpu-mediated proteasomal degradation of the restriction factor. Cell Host Microbe 2009, 5, 285–297.
  40. Iwabu, Y.; Fujita, H.; Kinomoto, M.; Kaneko, K.; Ishizaka, Y.; Tanaka, Y.; Sata, T.; Tokunaga, K. HIV-1 accessory protein Vpu internalizes cell-surface BST-2/tetherin through transmembrane interactions leading to lysosomes. J. Biol. Chem. 2009, 284, 35060–35072.
  41. Mangeat, B.; Gers-Huber, G.; Lehmann, M.; Zufferey, M.; Luban, J.; Piguet, V. HIV-1 Vpu neutralizes the antiviral factor Tetherin/BST-2 by binding it and directing its beta-TrCP2-dependent degradation. PLoS Pathog. 2009, 5, e1000574.
  42. Skasko, M.; Wang, Y.; Tian, Y.; Tokarev, A.; Munguia, J.; Ruiz, A.; Stephens, E.B.; Opella, S.J.; Guatelli, J. HIV-1 Vpu protein antagonizes innate restriction factor BST-2 via lipid-embedded helix-helix interactions. J. Biol. Chem. 2012, 287, 58–67.
  43. Mitchell, R.S.; Katsura, C.; Skasko, M.A.; Fitzpatrick, K.; Lau, D.; Ruiz, A.; Stephens, E.B.; Margottin-Goguet, F.; Benarous, R.; Guatelli, J.C. Vpu antagonizes BST-2-mediated restriction of HIV-1 release via beta-TrCP and endo-lysosomal trafficking. PLoS Pathog. 2009, 5, e1000450.
  44. Liu, Y.; Song, Y.; Zhang, S.; Diao, M.; Huang, S.; Li, S.; Tan, X. PSGL-1 inhibits HIV-1 infection by restricting actin dynamics and sequestering HIV envelope proteins. Cell Discov. 2020, 6, 53.
  45. Goetz, D.J.; Greif, D.M.; Ding, H.; Camphausen, R.T.; Howes, S.; Comess, K.M.; Snapp, K.R.; Kansas, G.S.; Luscinskas, F.W. Isolated P-selectin glycoprotein ligand-1 dynamic adhesion to P- and E-selectin. J. Cell Biol. 1997, 137, 509–519.
  46. Xia, L.; Ramachandran, V.; McDaniel, J.M.; Nguyen, K.N.; Cummings, R.D.; McEver, R.P. N-terminal residues in murine P-selectin glycoprotein ligand-1 required for binding to murine P-selectin. Blood 2003, 101, 552–559.
  47. Spertini, O.; Cordey, A.S.; Monai, N.; Giuffre, L.; Schapira, M. P-selectin glycoprotein ligand 1 is a ligand for L-selectin on neutrophils, monocytes, and CD34+ hematopoietic progenitor cells. J. Cell Biol. 1996, 135, 523–531.
  48. Baisse, B.; Galisson, F.; Giraud, S.; Schapira, M.; Spertini, O. Evolutionary conservation of P-selectin glycoprotein ligand-1 primary structure and function. BMC Evol. Biol. 2007, 7, 166.
  49. Carlow, D.A.; Gossens, K.; Naus, S.; Veerman, K.M.; Seo, W.; Ziltener, H.J. PSGL-1 function in immunity and steady state homeostasis. Immunol. Rev. 2009, 230, 75–96.
  50. Moore, K.L. Structure and function of P-selectin glycoprotein ligand-1. Leuk Lymphoma 1998, 29, 1–15.
  51. Liu, Y.; Fu, Y.; Wang, Q.; Li, M.; Zhou, Z.; Dabbagh, D.; Fu, C.; Zhang, H.; Li, S.; Zhang, T.; et al. Proteomic profiling of HIV-1 infection of human CD4(+) T cells identifies PSGL-1 as an HIV restriction factor. Nat. Microbiol. 2019, 4, 813–825.
  52. Grover, J.R.; Veatch, S.L.; Ono, A. Basic motifs target PSGL-1, CD43, and CD44 to plasma membrane sites where HIV-1 assembles. J. Virol. 2015, 89, 454–467.
  53. Fu, Y.; He, S.; Waheed, A.A.; Dabbagh, D.; Zhou, Z.; Trinite, B.; Wang, Z.; Yu, J.; Wang, D.; Li, F.; et al. PSGL-1 restricts HIV-1 infectivity by blocking virus particle attachment to target cells. Proc. Natl. Acad. Sci. USA 2020, 117, 9537–9545.
  54. Aggarwal, A.; Stella, A.O.; Henry, C.C.; Narayan, K.; Turville, S.G. Embedding of HIV Egress within Cortical F-Actin. Pathogens 2022, 11, 56.
  55. Weissbach, L.; Settleman, J.; Kalady, M.F.; Snijders, A.J.; Murthy, A.E.; Yan, Y.X.; Bernards, A. Identification of a human rasGAP-related protein containing calmodulin-binding motifs. J. Biol. Chem. 1994, 269, 20517–20521.
  56. Bashour, A.M.; Fullerton, A.T.; Hart, M.J.; Bloom, G.S. IQGAP1, a Rac- and Cdc42-binding protein, directly binds and cross-links microfilaments. J. Cell Biol. 1997, 137, 1555–1566.
  57. Liu, B.; Dai, R.; Tian, C.J.; Dawson, L.; Gorelick, R.; Yu, X.F. Interaction of the human immunodeficiency virus type 1 nucleocapsid with actin. J. Virol. 1999, 73, 2901–2908.
  58. Poole, E.; Strappe, P.; Mok, H.P.; Hicks, R.; Lever, A.M. HIV-1 Gag-RNA interaction occurs at a perinuclear/centrosomal site; analysis by confocal microscopy and FRET. Traffic 2005, 6, 741–755.
  59. Gladnikoff, M.; Shimoni, E.; Gov, N.S.; Rousso, I. Retroviral assembly and budding occur through an actin-driven mechanism. Biophys. J. 2009, 97, 2419–2428.
  60. Carlson, L.A.; de Marco, A.; Oberwinkler, H.; Habermann, A.; Briggs, J.A.; Krausslich, H.G.; Grunewald, K. Cryo electron tomography of native HIV-1 budding sites. PLoS Pathog. 2010, 6, e1001173.
  61. Stauffer, S.; Rahman, S.A.; de Marco, A.; Carlson, L.A.; Glass, B.; Oberwinkler, H.; Herold, N.; Briggs, J.A.; Muller, B.; Grunewald, K.; et al. The nucleocapsid domain of Gag is dispensable for actin incorporation into HIV-1 and for association of viral budding sites with cortical F-actin. J. Virol. 2014, 88, 7893–7903.
  62. Rahman, S.A.; Koch, P.; Weichsel, J.; Godinez, W.J.; Schwarz, U.; Rohr, K.; Lamb, D.C.; Krausslich, H.G.; Muller, B. Investigating the role of F-actin in human immunodeficiency virus assembly by live-cell microscopy. J. Virol. 2014, 88, 7904–7914.
  63. Valenzuela-Fernandez, A.; Alvarez, S.; Gordon-Alonso, M.; Barrero, M.; Ursa, A.; Cabrero, J.R.; Fernandez, G.; Naranjo-Suarez, S.; Yanez-Mo, M.; Serrador, J.M.; et al. Histone deacetylase 6 regulates human immunodeficiency virus type 1 infection. Mol. Biol. Cell 2005, 16, 5445–5454.
  64. Valenzuela-Fernandez, A.; Cabrero, J.R.; Serrador, J.M.; Sanchez-Madrid, F. HDAC6: A key regulator of cytoskeleton, cell migration and cell-cell interactions. Trends Cell Biol. 2008, 18, 291–297.
  65. Sattentau, Q. Avoiding the void: Cell-to-cell spread of human viruses. Nat. Rev. Microbiol. 2008, 6, 815–826.
  66. Vasiliver-Shamis, G.; Cho, M.W.; Hioe, C.E.; Dustin, M.L. Human immunodeficiency virus type 1 envelope gp120-induced partial T-cell receptor signaling creates an F-actin-depleted zone in the virological synapse. J. Virol. 2009, 83, 11341–11355.
  67. Sol-Foulon, N.; Sourisseau, M.; Porrot, F.; Thoulouze, M.I.; Trouillet, C.; Nobile, C.; Blanchet, F.; di Bartolo, V.; Noraz, N.; Taylor, N.; et al. ZAP-70 kinase regulates HIV cell-to-cell spread and virological synapse formation. EMBO J. 2007, 26, 516–526.
  68. Cabrera-Rodriguez, R.; Perez-Yanes, S.; Lorenzo-Sanchez, I.; Estevez-Herrera, J.; Garcia-Luis, J.; Trujillo-Gonzalez, R.; Valenzuela-Fernandez, A. TDP-43 Controls HIV-1 Viral Production and Virus Infectiveness. Int. J. Mol. Sci. 2023, 24, 7658.
  69. Marrero-Hernandez, S.; Marquez-Arce, D.; Cabrera-Rodriguez, R.; Estevez-Herrera, J.; Perez-Yanes, S.; Barroso-Gonzalez, J.; Madrid, R.; Machado, J.D.; Blanco, J.; Valenzuela-Fernandez, A. HIV-1 Nef Targets HDAC6 to Assure Viral Production and Virus Infection. Front. Microbiol. 2019, 10, 2437.
  70. Valera, M.S.; de Armas-Rillo, L.; Barroso-Gonzalez, J.; Ziglio, S.; Batisse, J.; Dubois, N.; Marrero-Hernandez, S.; Borel, S.; Garcia-Exposito, L.; Biard-Piechaczyk, M.; et al. The HDAC6/APOBEC3G complex regulates HIV-1 infectiveness by inducing Vif autophagic degradation. Retrovirology 2015, 12, 53.
  71. Nishi, M.; Ryo, A.; Tsurutani, N.; Ohba, K.; Sawasaki, T.; Morishita, R.; Perrem, K.; Aoki, I.; Morikawa, Y.; Yamamoto, N. Requirement for microtubule integrity in the SOCS1-mediated intracellular dynamics of HIV-1 Gag. FEBS Lett. 2009, 583, 1243–1250.
  72. Yoshimura, A.; Naka, T.; Kubo, M. SOCS proteins, cytokine signalling and immune regulation. Nat. Rev. Immunol. 2007, 7, 454–465.
  73. Ryo, A.; Tsurutani, N.; Ohba, K.; Kimura, R.; Komano, J.; Nishi, M.; Soeda, H.; Hattori, S.; Perrem, K.; Yamamoto, M.; et al. SOCS1 is an inducible host factor during HIV-1 infection and regulates the intracellular trafficking and stability of HIV-1 Gag. Proc. Natl. Acad. Sci. USA 2008, 105, 294–299.
  74. Briggs, M.W.; Sacks, D.B. IQGAP proteins are integral components of cytoskeletal regulation. EMBO Rep. 2003, 4, 571–574.
  75. Choi, S.; Anderson, R.A. IQGAP1 is a phosphoinositide effector and kinase scaffold. Adv. Biol. Regul. 2016, 60, 29–35.
  76. Hedman, A.C.; Smith, J.M.; Sacks, D.B. The biology of IQGAP proteins: Beyond the cytoskeleton. EMBO Rep. 2015, 16, 427–446.
  77. Sabo, Y.; de Los Santos, K.; Goff, S.P. IQGAP1 Negatively Regulates HIV-1 Gag Trafficking and Virion Production. Cell Rep. 2020, 30, 4065–4081.e4.
  78. Dolnik, O.; Kolesnikova, L.; Welsch, S.; Strecker, T.; Schudt, G.; Becker, S. Interaction with Tsg101 is necessary for the efficient transport and release of nucleocapsids in marburg virus-infected cells. PLoS Pathog. 2014, 10, e1004463.
  79. Gladue, D.P.; Holinka, L.G.; Fernandez-Sainz, I.J.; Prarat, M.V.; O’Donnell, V.; Vepkhvadze, N.G.; Lu, Z.; Risatti, G.R.; Borca, M.V. Interaction between Core protein of classical swine fever virus with cellular IQGAP1 protein appears essential for virulence in swine. Virology 2011, 412, 68–74.
  80. Lu, J.; Qu, Y.; Liu, Y.; Jambusaria, R.; Han, Z.; Ruthel, G.; Freedman, B.D.; Harty, R.N. Host IQGAP1 and Ebola virus VP40 interactions facilitate virus-like particle egress. J. Virol. 2013, 87, 7777–7780.
  81. Chen, D.; Wang, M.; Zhou, S.; Zhou, Q. HIV-1 Tat targets microtubules to induce apoptosis, a process promoted by the pro-apoptotic Bcl-2 relative Bim. EMBO J. 2002, 21, 6801–6810.
  82. Dumas, A.; Le-Bury, G.; Marie-Anais, F.; Herit, F.; Mazzolini, J.; Guilbert, T.; Bourdoncle, P.; Russell, D.G.; Benichou, S.; Zahraoui, A.; et al. The HIV-1 protein Vpr impairs phagosome maturation by controlling microtubule-dependent trafficking. J. Cell Biol. 2015, 211, 359–372.
  83. Stewart, S.A.; Poon, B.; Song, J.Y.; Chen, I.S. Human immunodeficiency virus type 1 vpr induces apoptosis through caspase activation. J. Virol. 2000, 74, 3105–3111.
  84. Muthumani, K.; Hwang, D.S.; Desai, B.M.; Zhang, D.; Dayes, N.; Green, D.R.; Weiner, D.B. HIV-1 Vpr induces apoptosis through caspase 9 in T cells and peripheral blood mononuclear cells. J. Biol. Chem. 2002, 277, 37820–37831.
  85. Moon, H.S.; Yang, J.S. Role of HIV Vpr as a regulator of apoptosis and an effector on bystander cells. Mol. Cells 2006, 21, 7–20.
  86. Creagh, E.M.; Conroy, H.; Martin, S.J. Caspase-activation pathways in apoptosis and immunity. Immunol. Rev. 2003, 193, 10–21.
  87. Fan, T.J.; Han, L.H.; Cong, R.S.; Liang, J. Caspase family proteases and apoptosis. Acta Biochim. Biophys. Sin. 2005, 37, 719–727.
  88. Green, D.R. Caspase Activation and Inhibition. Cold Spring Harb. Perspect Biol. 2022, 14, a041020.
  89. Davis, M.J.; Swanson, J.A. Technical advance: Caspase-1 activation and IL-1beta release correlate with the degree of lysosome damage, as illustrated by a novel imaging method to quantify phagolysosome damage. J. Leukoc. Biol. 2010, 88, 813–822.
  90. Danielsson, F.; Peterson, M.K.; Caldeira Araujo, H.; Lautenschlager, F.; Gad, A.K.B. Vimentin Diversity in Health and Disease. Cells 2018, 7, 147.
  91. Nekrasova, O.E.; Mendez, M.G.; Chernoivanenko, I.S.; Tyurin-Kuzmin, P.A.; Kuczmarski, E.R.; Gelfand, V.I.; Goldman, R.D.; Minin, A.A. Vimentin intermediate filaments modulate the motility of mitochondria. Mol. Biol. Cell 2011, 22, 2282–2289.
  92. Matveeva, E.A.; Venkova, L.S.; Chernoivanenko, I.S.; Minin, A.A. Vimentin is involved in regulation of mitochondrial motility and membrane potential by Rac1. Biol. Open 2015, 4, 1290–1297.
  93. Hirata, Y.; Hosaka, T.; Iwata, T.; Le, C.T.; Jambaldorj, B.; Teshigawara, K.; Harada, N.; Sakaue, H.; Sakai, T.; Yoshimoto, K.; et al. Vimentin binds IRAP and is involved in GLUT4 vesicle trafficking. Biochem. Biophys. Res. Commun. 2011, 405, 96–101.
  94. Margiotta, A.; Bucci, C. Role of Intermediate Filaments in Vesicular Traffic. Cells 2016, 5, 20.
  95. Styers, M.L.; Salazar, G.; Love, R.; Peden, A.A.; Kowalczyk, A.P.; Faundez, V. The endo-lysosomal sorting machinery interacts with the intermediate filament cytoskeleton. Mol. Biol. Cell 2004, 15, 5369–5382.
  96. Ikawa, K.; Satou, A.; Fukuhara, M.; Matsumura, S.; Sugiyama, N.; Goto, H.; Fukuda, M.; Inagaki, M.; Ishihama, Y.; Toyoshima, F. Inhibition of endocytic vesicle fusion by Plk1-mediated phosphorylation of vimentin during mitosis. Cell Cycle 2014, 13, 126–137.
  97. Potokar, M.; Kreft, M.; Li, L.; Daniel Andersson, J.; Pangrsic, T.; Chowdhury, H.H.; Pekny, M.; Zorec, R. Cytoskeleton and vesicle mobility in astrocytes. Traffic 2007, 8, 12–20.
  98. Ivaska, J.; Vuoriluoto, K.; Huovinen, T.; Izawa, I.; Inagaki, M.; Parker, P.J. PKCepsilon-mediated phosphorylation of vimentin controls integrin recycling and motility. EMBO J. 2005, 24, 3834–3845.
  99. Vardjan, N.; Gabrijel, M.; Potokar, M.; Svajger, U.; Kreft, M.; Jeras, M.; de Pablo, Y.; Faiz, M.; Pekny, M.; Zorec, R. IFN-gamma-induced increase in the mobility of MHC class II compartments in astrocytes depends on intermediate filaments. J. Neuroinflammation 2012, 9, 144.
  100. Potokar, M.; Stenovec, M.; Gabrijel, M.; Li, L.; Kreft, M.; Grilc, S.; Pekny, M.; Zorec, R. Intermediate filaments attenuate stimulation-dependent mobility of endosomes/lysosomes in astrocytes. Glia 2010, 58, 1208–1219.
  101. Zhang, Y.; Wen, Z.; Shi, X.; Liu, Y.J.; Eriksson, J.E.; Jiu, Y. The diverse roles and dynamic rearrangement of vimentin during viral infection. J. Cell Sci. 2020, 134, jcs250597.
  102. Ramos, I.; Stamatakis, K.; Oeste, C.L.; Perez-Sala, D. Vimentin as a Multifaceted Player and Potential Therapeutic Target in Viral Infections. Int. J. Mol. Sci. 2020, 21, 4675.
  103. Helfand, B.T.; Mikami, A.; Vallee, R.B.; Goldman, R.D. A requirement for cytoplasmic dynein and dynactin in intermediate filament network assembly and organization. J. Cell Biol. 2002, 157, 795–806.
  104. Gyoeva, F.K.; Gelfand, V.I. Coalignment of vimentin intermediate filaments with microtubules depends on kinesin. Nature 1991, 353, 445–448.
  105. Yoon, M.; Moir, R.D.; Prahlad, V.; Goldman, R.D. Motile properties of vimentin intermediate filament networks in living cells. J. Cell Biol. 1998, 143, 147–157.
  106. Prahlad, V.; Yoon, M.; Moir, R.D.; Vale, R.D.; Goldman, R.D. Rapid movements of vimentin on microtubule tracks: Kinesin-dependent assembly of intermediate filament networks. J. Cell Biol. 1998, 143, 159–170.
  107. Liao, G.; Gundersen, G.G. Kinesin is a candidate for cross-bridging microtubules and intermediate filaments. Selective binding of kinesin to detyrosinated tubulin and vimentin. J. Biol. Chem. 1998, 273, 9797–9803.
  108. Robert, A.; Herrmann, H.; Davidson, M.W.; Gelfand, V.I. Microtubule-dependent transport of vimentin filament precursors is regulated by actin and by the concerted action of Rho- and p21-activated kinases. FASEB J. 2014, 28, 2879–2890.
  109. Hookway, C.; Ding, L.; Davidson, M.W.; Rappoport, J.Z.; Danuser, G.; Gelfand, V.I. Microtubule-dependent transport and dynamics of vimentin intermediate filaments. Mol. Biol. Cell 2015, 26, 1675–1686.
  110. Clarke, E.J.; Allan, V. Intermediate filaments: Vimentin moves in. Curr. Biol. 2002, 12, R596–R598.
  111. Aloisi, A.L.; Bucci, C. Rab GTPases-cargo direct interactions: Fine modulators of intracellular trafficking. Histol. Histopathol. 2013, 28, 839–849.
  112. Guerra, F.; Bucci, C. Multiple Roles of the Small GTPase Rab7. Cells 2016, 5, 34.
  113. Cogli, L.; Progida, C.; Bramato, R.; Bucci, C. Vimentin phosphorylation and assembly are regulated by the small GTPase Rab7a. Biochim. Biophys. Acta 2013, 1833, 1283–1293.
  114. Margiotta, A.; Progida, C.; Bakke, O.; Bucci, C. Rab7a regulates cell migration through Rac1 and vimentin. Biochim. Biophys. Acta Mol. Cell Res 2017, 1864, 367–381.
  115. Romano, R.; Del Fiore, V.S.; Bucci, C. Role of the Intermediate Filament Protein Peripherin in Health and Disease. Int. J. Mol. Sci. 2022, 23, 15416.
  116. Fay, N.; Pante, N. The intermediate filament network protein, vimentin, is required for parvoviral infection. Virology 2013, 444, 181–190.
  117. Risco, C.; Rodriguez, J.R.; Lopez-Iglesias, C.; Carrascosa, J.L.; Esteban, M.; Rodriguez, D. Endoplasmic reticulum-Golgi intermediate compartment membranes and vimentin filaments participate in vaccinia virus assembly. J. Virol. 2002, 76, 1839–1855.
  118. Sabharwal, P.; Amritha, C.K.; Sushmitha, C.; Natraj, U.; Savithri, H.S. Intracellular trafficking and endocytic uptake pathway of Pepper vein banding virus-like particles in epithelial cells. Nanomedicine 2019, 14, 1247–1265.
  119. Wu, W.; Pante, N. Vimentin plays a role in the release of the influenza A viral genome from endosomes. Virology 2016, 497, 41–52.
  120. Fernandez-Ortega, C.; Ramirez, A.; Casillas, D.; Paneque, T.; Ubieta, R.; Dubed, M.; Navea, L.; Castellanos-Serra, L.; Duarte, C.; Falcon, V.; et al. Identification of Vimentin as a Potential Therapeutic Target against HIV Infection. Viruses 2016, 8, 98.
  121. Koths, K.; Taylor, E.; Halenbeck, R.; Casipit, C.; Wang, A. Cloning and characterization of a human Mac-2-binding protein, a new member of the superfamily defined by the macrophage scavenger receptor cysteine-rich domain. J. Biol. Chem. 1993, 268, 14245–14249.
  122. Groschel, B.; Braner, J.J.; Funk, M.; Linde, R.; Doerr, H.W.; Cinatl, J., Jr.; Iacobelli, S. Elevated plasma levels of 90K (Mac-2 BP) immunostimulatory glycoprotein in HIV-1-infected children. J. Clin. Immunol. 2000, 20, 117–122.
  123. Tinari, N.; Natoli, C.; D’Ostilio, N.; Ghinelli, F.; Sighinolfi, L.; Ortona, L.; Tamburrini, E.; Piazza, M.; Chirianni, A.; Guerra, L.; et al. 90K (Mac-2 BP) predicts CD4 decline in human immunodeficiency virus-infected patients with CD4 counts above 200 × 10(6) cells/L. Arch. Pathol. Lab. Med. 1998, 122, 178–181.
  124. Darcissac, E.C.; Vidal, V.; De La Tribonniere, X.; Mouton, Y.; Bahr, G.M. Variations in serum IL-7 and 90K/Mac-2 binding protein (Mac-2 BP) levels analysed in cohorts of HIV-1 patients and correlated with clinical changes following antiretroviral therapy. Clin. Exp. Immunol. 2001, 126, 287–294.
  125. Wang, Q.; Zhang, X.; Han, Y.; Wang, X.; Gao, G. M2BP inhibits HIV-1 virion production in a vimentin filaments-dependent manner. Sci. Rep. 2016, 6, 32736.
  126. Adamson, C.S.; Freed, E.O. Human immunodeficiency virus type 1 assembly, release, and maturation. Adv. Pharmacol. 2007, 55, 347–387.
  127. Jia, X.; Liu, L.; Liu, X.; Wu, D.; Yin, L.; Liu, X.; Zhang, J.; Yang, P.; Lu, H.; Zhang, L. Vimentin-a potential biomarker for therapeutic efficiency of HAART. Acta Biochim. Biophys. Sin. 2014, 46, 1001–1006.
  128. Shoeman, R.L.; Honer, B.; Stoller, T.J.; Kesselmeier, C.; Miedel, M.C.; Traub, P.; Graves, M.C. Human immunodeficiency virus type 1 protease cleaves the intermediate filament proteins vimentin, desmin, and glial fibrillary acidic protein. Proc. Natl. Acad. Sci. USA 1990, 87, 6336–6340.
  129. Shoeman, R.L.; Mothes, E.; Honer, B.; Kesselmeier, C.; Traub, P. Effect of human immunodeficiency virus type 1 protease on the intermediate filament subunit protein vimentin: Cleavage, in vitro assembly and altered distribution of filaments in vivo following microinjection of protease. Acta Histochem. Suppl. 1991, 41, 129–141.
  130. Shoeman, R.L.; Mothes, E.; Kesselmeier, C.; Traub, P. Intermediate filament assembly and stability in vitro: Effect and implications of the removal of head and tail domains of vimentin by the human immunodeficiency virus type 1 protease. Cell Biol. Int. Rep. 1990, 14, 583–594.
  131. Karczewski, M.K.; Strebel, K. Cytoskeleton association and virion incorporation of the human immunodeficiency virus type 1 Vif protein. J. Virol. 1996, 70, 494–507.
  132. Shoeman, R.L.; Honer, B.; Mothes, E.; Traub, P. Potential role of the viral protease in human immunodeficiency virus type 1 associated pathogenesis. Med. Hypotheses 1992, 37, 137–150.
  133. Armenian, H.K.; Hoover, D.R.; Rubb, S.; Metz, S.; Martinez-Maza, O.; Chmiel, J.; Kingsley, L.; Saah, A. Risk factors for non-Hodgkin’s lymphomas in acquired immunodeficiency syndrome (AIDS). Am. J. Epidemiol 1996, 143, 374–379.
  134. Chou, K.C.; Tomasselli, A.G.; Reardon, I.M.; Heinrikson, R.L. Predicting human immunodeficiency virus protease cleavage sites in proteins by a discriminant function method. Proteins 1996, 24, 51–72.
More
This entry is offline, you can click here to edit this entry!
Video Production Service