Literature
首页医源资料库在线期刊美国生理学杂志2006年第289卷第2期

Persistent NF-B activation in renal epithelial cells in a mouse model of HIV-associated nephropathy

来源:美国生理学杂志
摘要:【关键词】cellsDepartmentofMedicineandRammelkampCenterforEducationandResearch,MetroHealthMedicalCenter,CaseWesternReserveUniversitySchoolofMedicine,Cleveland,OhioABSTRACTHumanimmunodeficiencyvirus(HIV)-associatednephropathy(HIVAN)iscaused,inpart,bydirectin......

点击显示 收起

【关键词】  cells

    Department of Medicine and Rammelkamp Center for Education and Research, MetroHealth Medical Center, Case Western Reserve University School of Medicine, Cleveland, Ohio

    ABSTRACT

    Human immunodeficiency virus (HIV)-associated nephropathy (HIVAN) is caused, in part, by direct infection of kidney epithelial cells by HIV-1. In the spectrum of pathogenic host-virus interactions, abnormal activation or suppression of host transcription factors is common. NF-B is a necessary host transcription factor for HIV-1 gene expression, and it has been shown that NF-B activity is dysregulated in many naturally infected cell types. We show here that renal glomerular epithelial cells (podocytes) expressing the HIV-1 genome, similar to infected immune cells, also have a dysregulated and persistent activation of NF-B. Although podocytes produce p50, p52, RelA, RelB, and c-Rel, electrophoretic mobility shift assays and immunocytochemistry showed a predominant nuclear accumulation of p50/RelA-containing NF-B dimers in HIV-1-expressing podocytes compared with normal. In addition, the expression level of a transfected NF-B reporter plasmid was significantly higher in HIVAN podocytes. The mechanism of NF-B activation involved increased phosphorylation of IB, resulting in an enhanced turnover of the IB protein. There was no evidence for regulation by IB or the alternate pathway of NF-B activation. Altered activation of this key host transcription factor likely plays a role in the well-described cellular phenotypic changes observed in HIVAN, such as proliferation. Studies with inhibitors of proliferation and NF-B suggest that NF-B activation may contribute to the proliferative mechanism in HIVAN. In addition, because NF-B regulates many aspects of inflammation, this dysregulation may also contribute to disease severity and progression through regulation of proinflammatory processes in the kidney microenvironment.

    podocyte; chronic renal disease; HIV-1; transcription

    HUMAN IMMUNODEFICIENCY VIRUS (HIV)-associated nephropathy (HIVAN) is a complication of HIV/AIDS that affects 1% of the seropositive population in the United States. It is a significant cause of chronic kidney disease and is currently the third leading cause of kidney failure in adult African American men (31, 45). A part of the disease process is the direct infection of kidney epithelial cells including the glomerular podocyte (7, 9, 29, 34, 42). This ultimately results in injury to the podocyte and glomerular damage characterized as collapsing focal segmental glomerulosclerosis (FSGS). Podocyte injury is frequently recognized as a central event in the initiation and progression of many forms of FSGS as the podocyte is the glomerular cell type primarily responsible for synthesizing the molecular components of the glomerular filtration barrier (3, 32).

    The specific host-virus interactions in renal cells that initiate pathology in HIVAN are not well understood. Because this renal disease has been successfully modeled in transgenic mice (11, 17, 23, 30) and rats (43), aspects of HIVAN pathogenesis are likely to be independent of the infection process. In these instances, host cell pathogenesis may be caused by individual viral proteins, either intracellular or extracellular, which interfere with normal cellular function. Several HIV-1 proteins such as the regulatory proteins Tat, Nef, and Vpu and the envelope protein gp120 have been shown to participate in the pathogenesis of several HIV-related syndromes (52). In HIVAN, recent reports using a transgenic mouse model have proposed that the viral protein Nef may be responsible for many aspects of renal cell dysfunction (18, 19, 22, 53). These studies suggest that interference of normal cellular functions by individual viral proteins is central to HIVAN pathogenesis.

    A well-known example of viral interference in host cell function is the appropriation of host transcription factors to augment viral gene expression (21). In the HIV-1 life cycle, expression of the viral genome is dependent on the host transcriptional machinery, and HIV-1 has evolved to rely largely on NF-B (26). NF-B is a ubiquitous transcription factor that has a central role in mediating immune and inflammatory processes (8). The prototypic NF-B is a dimer consisting of a DNA-binding subunit (p50) and a trans-activating subunit (p65), which are regulated by a third protein, IB. There are multiple genes which code for the p50s (p105/p50 and p100/p52), p65s (RelA, RelB, and c-Rel), and IBs (IB, IB, IB, IB, Bcl-3), which result in combinatorial diversity of not only NF-B composition and variable DNA-binding specificities to B sites but also distinct activation mechanisms through interaction with the different IBs. NF-B is an immediate early response factor because the active heterodimer is preformed in the cytoplasm and is retained in the cytoplasm through interaction with an IB (Fig. 1). In the classic mechanism of activation, IB is phosphorylated by IB kinases (IKK), which target the IB for ubiquitination and proteasomal degradation. In the alternate mechanism of activation, the IB-like COOH terminus of p100/p52 is phosphorylated and degraded similar to an IB. In both pathways, this releases active heterodimers from their cytoplasmic anchors, exposing a nuclear localization signal, followed by nuclear translocation and binding to target gene promoters. Thus the key regulatory event in NF-B activation depends on the phosphorylation and subsequent degradation of an IB or the IB-like portion of p100.

    Previously, it has been shown that HIV-infected monocytes and macrophages have a modified NF-B activation cascade, causing a sustained enhancement of NF-B nuclear translocation (10, 35, 47). This persistent activation of NF-B supports the viral life cycle, but unfortunately, is also detrimental to normal host cell function. There are over 250 host genes known to be regulated by NF-B, and any dysregulation of NF-B can have a significant impact on normal cell behavior. We have previously shown that, similar to infected T cells, the expression of the HIV-1 genome in kidney cells is also dependent on NF-B (5). We now report here that, similar to infected immune cells, NF-B is persistently activated in epithelial cells from HIV-1 transgenic mouse kidneys. Using podocyte cell lines derived from normal and HIV-1 transgenic kidneys, we show that this persistent NF-B activation involved increased phosphorylation and turnover of IB, with a coincident increase in the nuclear appearance of p50/RelA heterodimers. There was no evidence for an IB-dependent activation mechanism or processing of p100 to suggest the involvement of the alternate pathway. Preliminary studies also indicated that HIV-induced NF-B activation was associated with increased proliferation in podocytes. In conjunction with our recent report of increased expression of other key NF-B target genes in HIVAN (46), these studies indicate that virus-induced activation of NF-B is an important cause of renal epithelial cell dysfunction in HIVAN pathogenesis.

    MATERIALS AND METHODS

    Cell lines and plasmids. The normal and HIV-1-expressing (HIVAN) podocyte cell lines were derived from an HIV-1 transgenic mouse model of HIVAN (30). This HIV-1 transgenic mouse model is a relevant model for the human disease process as the transgenic expression of the HIV-1 genome is similar to the integration and HIV-1 gene expression that occurs during the natural infection process. This small-animal model recreates virtually all aspects of the clinical course, pathology, and mortality of the human disease, and has been used extensively for molecular (1, 6, 33, 40, 41) and genetic (15) studies in host pathogenesis and for HIVAN therapy design and testing (4, 37, 50). Podocyte cell lines and growth conditions have been described and studies were performed with cells differentiated for 810 days under nonpermissive conditions (36, 48). An indicator plasmid for NF-B activity, pNF-B-SEAP (Clontech, Palo Alto, CA), expresses secreted alkaline phosphatase (SEAP) under the control of a promoter that contains four NF-B consensus sites. Control plasmids for transfection were a promoterless SEAP plasmid created by removing the 148-bp HindIII/BglII promoter region of pTAL-SEAP (Clontech) and a pCDNA3.1-based EGFP expression plasmid provided by J. Simske (MetroHealth Medical Center, Cleveland, OH). A dominant-negative mutant of IB, pCMV-IBM (Clontech), contains point mutations at serines 32 and 36 which block the phosphorylation events that trigger IB degradation and subsequent NF-B activation. Podocytes were transiently transfected using FuGENE 6 (Roche, Indianapolis, IN) at a 3:1 lipid-to-DNA ratio. Cells were plated in serum-containing media without antibiotics and fungicides the day before transfection in 35-mm dishes (105 cells/dish). Transfections were performed as recommended by the manufacturer with 1 μg of plasmid DNA and with continued culture in serum-containing media. Media and cells were assayed 48 h after transfection. Transfection efficiencies were determined as the percentage of EGFP-positive cells to total cells (DAPI-stained nuclei). Due to known differences in the growth rate of the HIVAN and normal podocytes (48), transfected cells were counted at the time of assay and data were normalized to cell number. SEAP activity was assayed from conditioned media using a chemiluminescence conversion kit (Great EscAPe, BD Biosciences) as directed. Luminescence was measured using a luminometer, and data are presented in relative light units (RLU) per 105 cells.

    Proliferation assay and inhibitors. Proliferation was suppressed using a double thymidine block or treatment with an agent that inhibits new DNA synthesis, 5-fluoruridine (5-FU). The thymidine block was performed using a standard procedure involving an 18-h incubation with 2 mM thymidine, release from arrest for 9 h, followed by a second block for 24 h. Cells were assessed at the end of the second block using the MTS assay (Promega, Madison, WI) according to manufacturer's instructions. Data are reported as absorbance (490 nm) minus background absorbance per 105 cells. A second method used a 24-h treatment with 5-FU (5 μM, Calbiochem) to starve cells for dTTP, blocking cells in the S-phase.

    Western blot analysis. Western blotting was performed on whole cell extracts using a standard RIPA protein extraction protocol (provided by the primary antibody supplier) containing protease inhibitors (mini-Complete, Roche). Proteins were resolved on 416% gradient denaturing gels and electroblotted to PVDF membranes. Rabbit polyclonal antibodies were purchased from Santa Cruz Biotechnology (Santa Cruz, CA) and include RelA p65 (c-20), RelB (c-19), c-Rel (c-21), NFKB1 p50 (c-19), NFKB2 p52 (k-27), IB (c-21), IB (c-20), IB (5177c), IB (m-121), and Bcl3 (c-14). The phospho-specific IB antibody was purchased from Cell Signaling Technologies (Beverly, MA), and the control antibody for -tubulin was from Sigma (St. Louis, MO). The secondary antibody (1:2,000 dilution) was a horseradish peroxidase-conjugated goat anti-rabbit (Jackson ImmunoResearch, West Grove, PA), and detection was by chemiluminescence (ECL, Amersham Biosciences, Piscataway, NJ) according to the manufacturer's instructions. The proteasome inhibitor MG132 (Calbiochem, San Diego, CA) was used at 10 μM for a 1.5-h pretreatment. The protein synthesis inhibitors anisomycin and emetine (Calbiochem) were used at 5 μM for a 16-h pretreatment. Phosphatase inhibitors (Protein Phosphatase Inhibitor Set, Calbiochem) were used at recommended concentrations.

    Immunohistochemistry. Cultured cells were grown on collagen-coated glass coverslips and were fixed in cold methanol. A monoclonal antibody against PCNA (Calbiochem) was used at 1:200 dilution. Secondary antibodies (1:400 dilution) were fluorescein isothiocyanate or cyanine 3-conjugated goat anti-rabbit or goat anti-mouse (Jackson ImmunoResearch), and coverslips were mounted with aqueous mounting media followed by epifluorescence microscopy.

    EMSA. Preparation of nuclear extracts, oligonucleotide labeling, binding reaction, and electrophoresis conditions have been described (5). The specificity of the shifted complexes was confirmed with a competition with 100-fold molar excess of nonradioactive NF-B oligonucleotide and also by supershifting DNA-protein complexes with the addition of anti-NF-B antibodies (0.5 μg) to the binding reaction. Initial experiments to identify gel migration positions of heterodimer and homodimer complexes were previously established with cell lines transfected with individual NF-B subunit proteins and have been published (5).

    Statistical analysis. Graphed data are representative experiments each conducted three times, in triplicate. Data are means ± SD with probability determined by t-test (2-tailed, 2-sample equal variance). P values used are provided in the figure legends. All EMSAs, Western blots, and immunohistochemical studies were performed, at minimum, three times from independent preparations of cells or protein extracts.

    RESULTS

    The composition of the NF-B complex and its regulation by the various IBs can define specific functions and mechanism of activation. Therefore, the expression pattern of all the NF-B and IB proteins was evaluated in podocytes. Mouse podocyte cell lines derived from normal and HIV-1 transgenic mice expressed all five of the NF-B proteins, RelA, RelB, c-Rel, p105/p50, and p100/p52 as determined by Western blotting (Fig. 2). The DNA-binding subunits p50 and p52 are synthesized as larger precursors and are posttranslationally processed by the 26S proteasome. The processing of p105 to p50 is a constitutive, cotranslational process and thus it is typical for the mature 50-kDa form to predominate. The processing of p100 to p52 is a highly regulated event involved in controlling NF-B activation via the alternate pathway (8). Thus in unstimulated cells it is typical to observe predominantly the full-length 100-kDa form. Overall, the expression pattern of these proteins was not different between normal and HIVAN podocytes (not shown). These cells also expressed two of the inhibitor proteins, IB and IB. There was no evidence that IB and IB were expressed by Western blotting (Fig. 2), and no expression of Bcl3 was observed by either Western blotting or immunocytochemistry (not shown).

    The Western blots indicated that podocytes express a variety of NF-B proteins; however, the NF-B transcription factor is functional only as a dimer. To determine which functional dimers form in podoctyes, EMSAs were performed using podocyte nuclear extracts and a consensus NF-B-binding site as a target DNA (Fig. 3A). EMSAs using normal podocyte nuclear extracts formed a single shift complex that migrated to a known position for p50 homodimers (see MATERIALS AND METHODS). These p50 homodimers are commonly found in both the cytoplasm and nucleus of most cells, and it is typical to observe DNA-binding p50 homodimers in unstimulated cells. However, the p50 homodimer does not contain a trans-activating subunit and therefore does not activate transcription. The EMSA using nuclear extracts from HIVAN podocytes resulted in two shifted bands; a less abundant band corresponding to p50 homodimers, and a second, more abundant and higher molecular weight complex that migrated at the location of a p50/p65 heterodimer. Thus the majority of shifting NF-B complexes in HIVAN podocytes were p50/p65 heterodimers, which would be typical of a cell responding to an NF-B stimulus.

    The identity of the DNA binding and trans-activating proteins that were present in the shift complexes was determined also using the EMSA. Antibodies specific for the individual NF-B proteins were added to the shift reactions (supershifts) before electrophoresis. Figure 3B shows the antibody supershifts using nuclear extracts prepared from HIVAN podocytes. The antibody to p50 recognized both the p50 homodimer and p50/p65 heterodimer, indicating that in both complexes, p50 was the DNA binding subunit. The antibody for RelA shifted the upper complex, whereas no supershifts were observed with RelB or c-Rel, indicating that the p50/p65 complex contained RelA. There was also no supershift with the p52 antibody. Thus the activated complex in the HIVAN podocytes was the prototypical p50/RelA heterodimer, which is typically the most abundant NF-B complex in most cells.

    To confirm the EMSA binding studies, we performed immunocytochemistry with antibodies against p50 and RelA comparing the expression and subcellular distribution pattern in normal and HIVAN podocyte (Fig. 3, C-F). Both p50 and RelA were easily detected in both HIVAN and normal podocytes at qualitatively similar levels of expression. The expression of p50 and RelA in the normal podocytes was more cytoplasmic in distribution, although faint staining was detected in the nucleus. In the HIVAN podocytes, the distribution of both p50 and RelA became more concentrated in the nucleus in most cells, indicating NF-B activation. Thus both the EMSA and immunocytochemistry studies suggest that enhanced NF-B activation of a p50/RelA complex is occurring in HIVAN podocytes.

    To extend these binding events to a functional assay, normal and HIVAN podocytes were transfected with an NF-B reporter plasmid, pNF-B-SEAP, to measure differences in NF-B activation between the two cell lines. Figure 4 is the quantification of marker gene (SEAP) expression in normal and HIVAN podocytes transfected with either the NF-B indicator SEAP plasmid or a control SEAP plasmid. There was a significantly higher level of NF-B activity in the HIVAN podocytes compared with normal. Cotransfection of the cells with an IB dominant negative efficiently blocked the NF-B-dependent reporter expression, reducing reporter expression to background levels. Transient transfection efficiencies for normal and HIVAN cells were similar with an average of 22.4 ± 2.7% (mean ± SD) for normal podocytes vs. 23.1 ± 1.8% for HIVAN podocytes. In addition, transfected cells were counted at the time of assay and data were normalized to cell number as there are known differences in the growth rate of the HIVAN and normal podocytes (48). The combination of this functional analysis and the in vitro binding evidence from the EMSA indicates that HIVAN podocytes had a higher level of NF-B activation compared with normal cells. Because no external stimulus was provided, this difference in NF-B activation was intrinsic to the HIVAN cells.

    Because the key step in NF-B activation is the phosphorylation and proteasomal degradation of IB, identifying the IB involved in the regulation of p50/RelA will be central to defining the mechanism of activation. Figure 2 indicated that podocytes express only IB and IB, and both these IBs are known to interact with p50/RelA. Western blotting was used to examine the expression pattern of the IBs in normal and HIVAN podocytes. In Fig. 5, A and B, two treatment strategies were used to evaluate IB turnover, including inhibition of proteasomal degradation with MG132 and inhibition of protein synthesis with emetine and anisomycin. These treatments were necessary as IB is rapidly degraded when phosphorylated, and also because NF-B activation increases new IB synthesis as the IB gene itself is regulated by NF-B. Thus the level of IB following an NF-B activation signal leads not only to increased IB degradation but also to new IB protein synthesis. First, the cells were treated with an inhibitor of the 26S proteasome, which prevents the degradation of the phosphorylated IB, but does not change NF-B activation. In these treatments, more IB was detected in HIVAN podocytes (Fig. 5A). In addition, reprobing this blot with an antibody specific for IB Ser-32 phosphorylation indicated that IB was phosphorylated to a higher degree in HIVAN podocytes compared with normal, which also is indicative of an NF-B activation event. The normal and HIVAN podocytes were next treated with protein synthesis inhibitors to permit visualization of IB degradation by the proteasome (Fig. 5B). The degradation of IB protein was more rapid in the HIVAN podocytes, also suggesting a higher level of NF-B activation.

    Although these studies strongly implied a role for IB, the expression of IB and p100 processing was also studied to exclude a possible contributing role of these two alternative regulatory mechanisms. Unlike IB, the IB gene is not regulated by NF-B, and thus NF-B activation results in a chronic downregulation of IB protein following a stimulus. In Fig. 5C, no differences in IB expression or chronic downregulation were observed between normal and HIVAN podocytes. Similarly, the inducible, proteasomal processing of p100 to p52 was also not different between normal and HIVAN podocytes, with both showing minimal processing of the p100 form (Fig. 5D). This would indicate that the alternate pathway was likely not participating in the observed NF-B activation. In summary, these studies indicated that the increased activation of p50/RelA in HIVAN podocytes was due to an increased phosphorylation and turnover of IB and did not appear to involve mechanisms involving IB or p100 processing.

    Because a central pathological finding in HIVAN is epithelial cell proliferation, we investigated a possible connection between podocyte proliferation and NF-B activation. In two sets of studies, NF-B activation was inhibited followed by an assay for proliferation, as well as the converse study, in which proliferation was inhibited and its effect on NF-B activity determined. In Fig. 6A, NF-B activity was significantly reduced by cotransfection with a dominant-negative inhibitor of IB (IBM). In this study, cell proliferation, measured concurrently with the NF-B assay, also was significantly reduced. In the converse experiment (Fig. 6B), however, inhibition of proliferation did not have a similar effect on NF-B activity. In this study, proliferation was reduced using two methods, a double thymidine block and treatment with 5-FU. Testing dose and treatment length variations of both blocking protocols were unable to fully inhibit proliferation in the podocytes; however, statistically significant reductions in proliferation were achieved, and the 5-FU treatment reduced proliferation to a level comparable to the level achieved with NF-B inhibition. The IC50 for 5-FU with regard to suppression of podocyte proliferation was determined using a standard dose-response curve (1 nM to 1 mM) and found to be 5 μM (data not shown). Treatments at the IC50 for 24 h did not result in significant cell death and therefore appeared to be below the apoptotic threshold for this treatment protocol. Although both treatments reduced proliferation, there was no similar reduction in NF-B activity in the HIVAN cells. In summary, it appeared that changes in the level of NF-B activation had an impact on proliferation, suggesting a possible stimulatory effect by NF-B. On the other hand, proliferation itself did not appear to be a driving force for NF-B activation and may suggest a possible sequence of events in which NF-B activation could be a contributing stimulus for proliferation in HIVAN.

    NF-B activation was also associated with the expression of proliferating cell nuclear antigen (PCNA), a marker for actively proliferating cells. As shown in Fig. 4, HIVAN cells exhibited greater nuclear accumulation of RelA (an indication of NF-B activation) compared with normal cells. In Fig. 7, PCNA localization was associated with cells that had nuclear staining for RelA, whereas cells without evidence of nuclear translocation of RelA were negative for PCNA (Fig. 7, arrow). This colocalization of PCNA-positive cells with nuclear RelA would suggest that NF-B activation is associated with actively proliferating cells.

    DISCUSSION

    The consequences of chronically dysregulated NF-B activation in renal epithelial cells likely contribute to HIVAN pathogenesis from not only enhancement of viral gene expression but also through the dysregulation of host genes. Previous studies and expression array screens comparing normal and transgenic mouse kidneys or cell lines have shown that a number of direct NF-B target genes are altered in HIVAN. These include the upregulation of cyclin D1, calcyclin, vascular cell adhesion molecule-1, intracellular adhesion molecule-1, fibronectin, vimentin, Fas, Fas ligand, and major histocompatibility class I heavy chain (37, 38, 44, 46). For some of these, connections could be inferred from the NF-B target gene and known pathogenic events in HIVAN. Our recent studies have shown a direct role for NF-B in regulating the expression of the key proapoptotic genes, Fas and Fas ligand, as a mechanism for renal epithelial apoptosis in HIVAN (46). As would be expected, some of these genes also are associated with normal immune responses. However, little mechanistic information is known about the contributing role of NF-B-regulated immune activation or inflammation to the overall pathogenic process in HIVAN.

    Indirect evidence exists that would suggest NF-B-regulated immune responses may have a contributing role in HIVAN pathogenesis. Before the era of highly active antiretroviral therapy, attempts to treat HIVAN patients employed corticosteroids with some success (12, 51, 54). An important part of the anti-inflammatory action of corticosteroids is through the suppression of NF-B (14). In these patients, improvements in renal function seemed to parallel the resolution of tubular interstitial damage and reduction in immune cell infiltrates. It is well known that HIV-1 infection is associated with a profound, systemic dysregulation of cytokines, and it has been proposed that a similar cytokine dysregulation may occur in the kidney microenvironment (27, 28). In addition, a recent study by Heckmann et al. (20) treated an HIV-1 transgenic mouse model with an inhibitor of IKK2, a kinase that phosphorylates the IBs. In this study, they found that treated animals had reduced proinflammatory cytokine production, reduced kidney immune cell infiltrates, and an overall improved renal pathology. The treatment, however, did not block the disease process entirely. This would be consistent with our earlier studies that demonstrated expression of HIV-1 genes in kidney cells was required for disease and that the renal disease could not be initiated by the dysregulated cytokine milieu alone (6). These observations in total suggest that direct HIV-1 infection and expression of viral proteins in renal cells are required for disease. It also suggests that NF-B activation, as a direct result of renal cell infection, may cause the renal expression of proinflammatory agents or may enhance immune cell recruitment and contribute to the overall disease severity or progression.

    The preliminary studies presented here would suggest that NF-B activation and proliferation in HIVAN are linked and that NF-B may be contributing to a proproliferative mechanism. NF-B activation and IB phosphorylation are hallmarks of diseases characterized by excessive proliferation such as chronic inflammation and cancer (13). NF-B is known to control both aspects of cell cycle progression as well as be stimulated by molecules active during the cell cycle, forming a necessary integration of gene transcription and the cell cycle (25, 39). One possible connection between NF-B activation and the proliferative defect in HIVAN may be through the transcriptional regulation of key cell cycle regulators such as D-type cyclins. Cyclin D1, as well as other cyclins and cyclin-dependent kinases, have been shown to be altered in HIVAN and the other collapsing glomerulopathies (2, 16, 38, 49). Cyclin D1 is known to be regulated by NF-B, and because cyclin D1 protein is degraded with each cell cycle, the level of cyclin D1 is largely controlled by new transcription (25). Additional studies would be needed to establish a direct relationship between NF-B activation and cyclin D1 expression in HIVAN.

    The contribution of an IB-regulated p50/RelA complex as the dysregulated NF-B in renal epithelial cells would be consistent with observations in other HIV-1-infected cell types, and thus this may be a common mechanism of viral interference in HIV-1-permissive cell types. Future studies will need to identify the upstream kinases that signal to the IKK complex. Identifying these upstream kinases also may lead to the identification of the viral protein(s) that induce the persistent NF-B activation, as there will likely be a direct link or interaction between a key signaling pathway and the viral protein. Recent studies have implicated Nef as candidate viral protein that participates in the proliferative and dedifferentiation phenotypes in HIVAN glomerular disease through activation of Src-kinase (19). Studies in other cell types have found that Nef also has a stimulatory effect on several transcription factors including NF-B, and thus Nef may also contribute to this aspect of renal cell dysfunction in HIVAN (24).

    In conclusion, we demonstrated that HIV-1 expression in renal epithelial cells causes a dysregulation in NF-B activation. This HIV-induced activation was dependent on IB phosphorylation and degradation, suggesting the involvement of the classic pathway of NF-B activation. This activation appeared to be linked to proliferation, a significant pathogenic phenotype of podocytes in HIVAN. Because NF-B has many host gene targets involved in mediating immune and inflammatory processes, this dysregulation will likely have many pathogenic effects such as proliferation and apoptosis (46) of infected renal cells, but also possibly adjacent, noninfected cells through the secretion of local-acting cytokines. Identifying both the mechanism of NF-B activation and its role in HIVAN pathogenesis may put forward new therapeutic strategies for HIVAN treatment. In addition, understanding NF-B's role in regulating inflammatory processes in the kidney also may have application to the management of other forms of chronic renal disease that include an inflammatory component.

    GRANTS

    This work was supported by National Institutes of Health Grant DK-61395.

    ACKNOWLEDGMENTS

    We thank Drs. J. Sedor and M. J. Ross for helpful comments and critical review of the manuscript and Dr. J. Simske for providing the EGFP expression plasmid.

    FOOTNOTES

    The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

    REFERENCES

    Barisoni L, Kriz W, Mundel P, and D'Agati V. The dysregulated podocyte phenotype: a novel concept in the pathogenesis of collapsing idiopathic focal segmental glomerulosclerosis and HIV-associated nephropathy. J Am Soc Nephrol 10: 5161, 1999.

    Barisoni L, Mokrzycki M, Sablay L, Nagata M, Yamase H, and Mundel P. Podocyte cell cycle regulation and proliferation in collapsing glomerulopathies. Kidney Int 58: 137143, 2000.

    Barisoni L and Mundel P. Podocyte biology and the emerging understanding of podocyte diseases. Am J Nephrol 23: 353360, 2003.

    Bird JE, Durham SK, Giancarli MR, Gitlitz PH, Pandya DG, Dambach DM, Mozes MM, and Kopp JB. Captopril prevents nephropathy in HIV-transgenic mice. J Am Soc Nephrol 9: 14411447, 1998.

    Bruggeman LA, Adler SH, and Klotman PE. Nuclear factor-B binding to the HIV-1 LTR in kidney: implications for HIV-associated nephropathy. Kidney Int 59: 21742181, 2001.

    Bruggeman LA, Dikman S, Meng C, Quaggin SE, Coffman TM, and Klotman PE. Nephropathy in human immunodeficiency virus-1 transgenic mice is due to renal transgene expression. J Clin Invest 100: 8492, 1997.

    Bruggeman LA, Ross MD, Tanji N, Cara A, Dikman S, Gordon RE, Burns GC, D'Agati VD, Winston JA, Klotman ME, and Klotman PE. Renal epithelium is a previously unrecognized site of HIV-1 infection. J Am Soc Nephrol 11: 20792087, 2000.

    Chen LF and Greene WC. Shaping the nuclear action of NF-B. Nat Rev Mol Cell Biol 5: 392401, 2004.

    Cohen AH, Sun NC, Shapshak P, and Imagawa DT. Demonstration of human immunodeficiency virus in renal epithelium in HIV-associated nephropathy. Mod Pathol 2: 125128, 1989.

    DeLuca C, Roulston A, Koromilas A, Wainberg MA, and Hiscott J. Chronic human immunodeficiency virus type 1 infection of myeloid cells disrupts the autoregulatory control of the NF-B/Rel pathway via enhanced IB degradation. J Virol 70: 51835193, 1996.

    Dickie P, Roberts A, Uwiera R, Witmer J, Sharma K, and Kopp JB. Focal glomerulosclerosis in proviral and c-fms transgenic mice links Vpr expression to HIV-associated nephropathy. Virology 322: 6981, 2004.

    Eustace JA, Nuermberger E, Choi M, Scheel PJ Jr, Moore R, and Briggs WA. Cohort study of the treatment of severe HIV-associated nephropathy with corticosteroids. Kidney Int 58: 12531260, 2000.

    Foo SY and Nolan GP. NF-B to the rescue: RELs, apoptosis and cellular transformation. Trends Genet 15: 229235, 1999.

    Funder JW. Glucocorticoid and mineralocorticoid receptors: biology and clinical relevance. Annu Rev Med 48: 231240, 1997.

    Gharavi AG, Ahmad T, Wong RD, Hooshyar R, Vaughn J, Oller S, Frankel RZ, Bruggeman LA, D'Agati VD, Klotman PE, and Lifton RP. Mapping a locus for susceptibility to HIV-1-associated nephropathy to mouse chromosome 3. Proc Natl Acad Sci USA 101: 24882493, 2004.

    Gherardi D, D'Agati V, Chu TH, Barnett A, Gianella-Borradori A, Gelman IH, and Nelson PJ. Reversal of collapsing glomerulopathy in mice with the cyclin-dependent kinase inhibitor CYC202. J Am Soc Nephrol 15: 12121222, 2004.

    Hanna Z, Kay DG, Cool M, Jothy S, Rebai N, and Jolicoeur P. Transgenic mice expressing human immunodeficiency virus type 1 in immune cells develop a severe AIDS-like disease. J Virol 72: 121132, 1998.

    Hanna Z, Kay DG, Rebai N, Guimond A, Jothy S, and Jolicoeur P. Nef harbors a major determinant of pathogenicity for an AIDS-like disease induced by HIV-1 in transgenic mice. Cell 95: 163175, 1998.

    He JC, Husain M, Sunamoto M, D'Agati VD, Klotman ME, Iyengar R, and Klotman PE. Nef stimulates proliferation of glomerular podocytes through activation of Src-dependent Stat3 and MAPK1,2 pathways. J Clin Invest 114: 643651, 2004.

    Heckmann A, Waltzinger C, Jolicoeur P, Dreano M, Kosco-Vilbois MH, and Sagot Y. IKK2 inhibitor alleviates kidney and wasting diseases in a murine model of human AIDS. Am J Pathol 164: 12531262, 2004.

    Hiscott J, Kwon H, and Genin P. Hostile takeovers: viral appropriation of the NF-B pathway. J Clin Invest 107: 143151, 2001.

    Husain M, Gusella GL, Klotman ME, Gelman IH, Ross MD, Schwartz EJ, Cara A, and Klotman PE. HIV-1 Nef induces proliferation and anchorage-independent growth in podocytes. J Am Soc Nephrol 13: 18061815, 2002.

    Jolicoeur P, Kay DG, Cool M, Jothy S, Rebai N, and Hanna Z. A novel mouse model of HIV-1 disease. Leukemia 13, Suppl 1: 7880, 1999.

    Joseph AM, Kumar M, and Mitra D. Nef: "necessary and enforcing factor" in HIV infection. Curr HIV Res 3: 8794, 2005.

    Joyce D, Albanese C, Steer J, Fu M, Bouzahzah B, and Pestell RG. NF-B and cell-cycle regulation: the cyclin connection. Cytokine Growth Factor Rev 12: 7390, 2001.

    Karn J. Tackling Tat. J Mol Biol 293: 235254, 1999.

    Kimmel PL. HIV-associated nephropathy: virologic issues related to renal sclerosis. Nephrol Dial Transplant 18, Suppl 6: vi59vi63, 2003.

    Kimmel PL, Cohen DJ, Abraham AA, Bodi I, Schwartz AM, and Phillips TM. Upregulation of MHC class II, interferon- and interferon- receptor protein expression in HIV-associated nephropathy. Nephrol Dial Transplant 18: 285292, 2003.

    Kimmel PL, Ferreira-Centeno A, Farkas-Szallasi T, Abraham AA, and Garrett CT. Viral DNA in microdissected renal biopsy tissue from HIV infected patients with nephrotic syndrome. Kidney Int 43: 13471352, 1993.

    Kopp JB, Klotman ME, Adler SH, Bruggeman LA, Dickie P, Marinos NJ, Eckhaus M, Bryant JL, Notkins AL, and Klotman PE. Progressive glomerulosclerosis and enhanced renal accumulation of basement membrane components in mice transgenic for human immunodeficiency virus type 1 genes. Proc Natl Acad Sci USA 89: 15771581, 1992.

    Kopp JB and Winkler C. HIV-associated nephropathy in African Americans. Kidney Int Suppl 83: S43S49, 2003.

    Kriz W and Lemley KV. The role of the podocyte in glomerulosclerosis. Curr Opin Nephrol Hypertens 8: 489497, 1999.

    Lewis W. Use of the transgenic mouse in models of AIDS cardiomyopathy. AIDS 17, Suppl 1: S36S45, 2003.

    Marras D, Bruggeman LA, Gao F, Tanji N, Mansukhani MM, Cara A, Ross MD, Gusella GL, Benson G, D'Agati VD, Hahn BH, Klotman ME, and Klotman PE. Replication and compartmentalization of HIV-1 in kidney epithelium of patients with HIV-associated nephropathy. Nat Med 8: 522526, 2002.

    McElhinny JA, MacMorran WS, Bren GD, Ten RM, Israel A, and Paya CV. Regulation of IB and p105 in monocytes and macrophages persistently infected with human immunodeficiency virus. J Virol 69: 15001509, 1995.

    Mundel P, Reiser J, Zuniga Mejia BA, Pavenstadt H, Davidson GR, Kriz W, and Zeller R. Rearrangements of the cytoskeleton and cell contacts induce process formation during differentiation of conditionally immortalized mouse podocyte cell lines. Exp Cell Res 236: 248258, 1997.

    Nelson PJ, D'Agati VD, Gries JM, Suarez JR, and Gelman IH. Amelioration of nephropathy in mice expressing HIV-1 genes by the cyclin-dependent kinase inhibitor flavopiridol. J Antimicrob Chemother 51: 921929, 2003.

    Nelson PJ, Sunamoto M, Husain M, and Gelman IH. HIV-1 expression induces cyclin D1 expression and pRb phosphorylation in infected podocytes: cell-cycle mechanisms contributing to the proliferative phenotype in HIV-associated nephropathy. BMC Microbiol 2: e26, 2002.

    Perkins ND, Felzien LK, Betts JC, Leung K, Beach DH, and Nabel GJ. Regulation of NF-B by cyclin-dependent kinases associated with the p300 coactivator. Science 275: 523527, 1997.

    Petermann A, Hiromura K, Pippin J, Blonski M, Couser WG, Kopp J, Mundel P, and Shankland SJ. Differential expression of d-type cyclins in podocytes in vitro and in vivo. Am J Pathol 164: 14171424, 2004.

    Ray PE, Bruggeman LA, Weeks BS, Kopp JB, Bryant JL, Owens JW, Notkins AL, and Klotman PE. bFGF and its low affinity receptors in the pathogenesis of HIV-associated nephropathy in transgenic mice. Kidney Int 46: 759772, 1994.

    Ray PE, Liu XH, Henry D III, Dye L, Xu L, Orenstein JM, and Schuztbank TE. Infection of human primary renal epithelial cells with HIV-1 from children with HIV-associated nephropathy. Kidney Int 53: 12171229, 1998.

    Ray PE, Liu XH, Robinson LR, Reid W, Xu L, Owens JW, Jones OD, Denaro F, Davis HG, and Bryant JL. A novel HIV-1 transgenic rat model of childhood HIV-1-associated nephropathy. Kidney Int 63: 22422253, 2003.

    Ross MD, Bruggeman LA, Hanss B, Sunamoto M, Marras D, Klotman ME, and Klotman PE. Podocan, a novel small leucine-rich repeat protein expressed in the sclerotic glomerular lesion of experimental HIV-associated nephropathy. J Biol Chem 278: 3324833255, 2003.

    Ross MJ and Klotman PE. HIV-associated nephropathy. AIDS 18: 10891099, 2004.

    Ross MJ, Martinka S, D'Agati VD, and Bruggeman LA. NF-B regulates Fas-mediated apoptosis in HIV-associated nephropathy. J Am Soc Nephrol 16: 24032411, 2005.

    Roulston A, Lin R, Beauparlant P, Wainberg MA, and Hiscott J. Regulation of human immunodeficiency virus type 1 and cytokine gene expression in myeloid cells by NF-B/Rel transcription factors. Microbiol Rev 59: 481505, 1995.

    Schwartz EJ, Cara A, Snoeck H, Ross MD, Sunamoto M, Reiser J, Mundel P, and Klotman PE. Human immunodeficiency virus-1 induces loss of contact inhibition in podocytes. J Am Soc Nephrol 12: 16771684, 2001.

    Shankland SJ, Eitner F, Hudkins KL, Goodpaster T, D'Agati V, and Alpers CE. Differential expression of cyclin-dependent kinase inhibitors in human glomerular disease: role in podocyte proliferation and maturation. Kidney Int 58: 674683, 2000.

    Shirai A and Klinman DM. Immunization with recombinant gp160 prolongs the survival of HIV-1 transgenic mice. AIDS Res Hum Retroviruses 9: 979983, 1993.

    Smith MC, Austen JL, Carey JT, Emancipator SN, Herbener T, Gripshover B, Mbanefo C, Phinney M, Rahman M, Salata RA, Weigel K, and Kalayjian RC. Prednisone improves renal function and proteinuria in human immunodeficiency virus-associated nephropathy. Am J Med 101: 4148, 1996.

    Stevenson M. HIV-1 pathogenesis. Nat Med 9: 853860, 2003.

    Sunamoto M, Husain M, He JC, Schwartz EJ, and Klotman PE. Critical role for Nef in HIV-1-induced podocyte dedifferentiation. Kidney Int 64: 16951701, 2003.

    Winston JA, Burns GC, and Klotman PE. Treatment of HIV-associated nephropathy. Semin Nephrol 20: 293298, 2000.

作者: Scott Martinka and Leslie A. Bruggeman 2013-9-26
医学百科App—中西医基础知识学习工具
  • 相关内容
  • 近期更新
  • 热文榜
  • 医学百科App—健康测试工具