ABSTRACT
Many viruses, including mammarenaviruses, have evolved mechanisms to counteract different components of the host cell innate immunity, which is required to facilitate robust virus multiplication. The double-stranded RNA (dsRNA) sensor protein kinase receptor (PKR) pathway plays a critical role in the cell anti-viral response. Whether PKR can restrict the multiplication of the Old World mammarenavirus lymphocytic choriomeningitis virus (LCMV) and the mechanisms by which LCMV may counteract the anti-viral functions of PKR have not yet been investigated. Here we present evidence that LCMV infection results in very limited levels of PKR activation, but LCMV multiplication is enhanced in the absence of PKR. In contrast, infection with a recombinant LCMV with a mutation affecting the 3′−5′ exonuclease (ExoN) activity of the viral nucleoprotein resulted in robust PKR activation in the absence of detectable levels of dsRNA, which was associated with severely restricted virus multiplication that was alleviated in the absence of PKR. However, pharmacological inhibition of PKR activation resulted in reduced levels of LCMV multiplication. These findings uncovered a complex role of the PKR pathway in LCMV-infected cells involving both pro- and anti-viral activities.
IMPORTANCE
As with many other viruses, the prototypic Old World mammarenavirus LCMV can interfere with the host cell innate immune response to infection, which includes the dsRNA sensor PKR pathway. A detailed understanding of LCMV-PKR interactions can provide novel insights about mammarenavirus-host cell interactions and facilitate the development of effective anti-viral strategies against human pathogenic mammarenaviruses. In the present work, we present evidence that LCMV multiplication is enhanced in PKR-deficient cells, but pharmacological inhibition of PKR activation unexpectedly resulted in severely restricted propagation of LCMV. Likewise, we document a robust PKR activation in LCMV-infected cells in the absence of detectable levels of dsRNA. Our findings have revealed a complex role of the PKR pathway during LCMV infection and uncovered the activation of PKR as a druggable target for the development of anti-viral drugs against human pathogenic mammarenaviruses.
KEYWORDS: mammarenaviruses, LCMV, PKR pathway, anti-viral, eIF2a, dsRNA, MAVS, RNase L, interferon, 3′–5′ exonuclease
INTRODUCTION
Mammarenaviruses are enveloped viruses with a bi-segmented negative-stranded RNA genome (1) which cause chronic infections of their natural rodent host across the world, and human infections occur through mucosal exposure to aerosols or by direct contact of abraded skin with infectious materials (2). Upon a zoonotic event, mammarenaviruses can subvert the innate immune responses in the infected individual and interfere with the development of an effective adaptive immune response, which results in unrestricted virus multiplication and associated pathology and disease. Thus, several mammarenaviruses, chiefly Lassa virus (LASV), cause hemorrhagic fever disease in humans and pose important public health problems in their endemic regions (3). Moreover, evidence indicates that the worldwide distributed lymphocytic choriomeningitis virus (LCMV) is a neglected human pathogen of clinical significance (4–8) and a serious risk to immunocompromised individuals (9, 10). There are no Food and Drug Administration-licensed arenavirus vaccines, and current anti-arenaviral therapy is limited to an off-label use of ribavirin for which efficacy remains controversial (11). Therefore, there is a pressing need to develop effective strategies to combat human pathogenic mammarenaviruses, a task that will be facilitated by a better understanding of mammarenavirus-host innate defense interactions.
During multiplication in their host cells, RNA viruses generate a variety of pathogen-associated molecular patterns (PAMPs), including RNA-based PAMPs, that are recognized by host cellular pathogen recognition receptor (PRR) molecules, including TLR 3/TLR 7 and the retinoic acid-inducible gene I-like receptors (RLRs) retinoic acid-inducible gene I (RIG-I) and MDA5 (12, 13). Activated RLRs associate with the adapter mitochondrial anti-viral-signaling (MAVS) protein (14) to promote activation of the non-classical IkB kinase (IKK)-related kinases (15) TANK-binding kinase 1 and IKK epsilon (16, 17) that activate interferon regulatory factor 3 and nuclear factor kappa-light-chain-enhancer of activated B cells, which, together with ATF2/c-JUN, initiate transcription of interferon beta (IFN-β) promoter (18). Interaction of secreted IFN-β with its receptor (IFNAR) activates the JAK/STAT signaling pathway (19), resulting in induction of hundreds of type 1 IFN (T1IFN) stimulated genes to produce a cellular anti-viral state and control viral infection. Many viruses, including mammarenaviruses (20), have developed ways to subvert the T1IFN response (21, 22). Induction of IFN-β production in mammarenavirus-infected cells is greatly diminished by the virus nucleoprotein (NP) ability to inhibit activation of IRF3 (23, 24). The anti-IFN-β activity of NP has been linked to its C-terminally located functional 3′−5′ exonuclease (ExoN) domain characteristic of the DEDDh ExoN superfamily (25, 26). It is thought that NP’s ExoN activity promotes the degradation of viral RNA species that can activate the RIG-I/MAVS pathway and subsequent induction of IFN-β production to trigger the T1IFN pathway (27). The double-stranded RNA (dsRNA) sensor protein kinase receptor (PKR) is at the center of cellular responses to a variety of stress signals, including viral infection (28). Short viral dsRNA species generated during virus replication in infected cells can trigger PKR activation reflected in increased levels of phosphorylated protein kinase receptor (pPKR) (29–32). Increased levels of pPKR lead to phosphorylation of the eukaryotic translation initiation factor 2 alpha (eIF2α), resulting in the inhibition of cap-dependent protein translation initiation, which contributes to restricting virus propagation (33). Accordingly, many viruses have evolved mechanisms to counteract the activation of the PKR pathway (34–36). However, conflicting findings have been reported on mammarenavirus-PKR pathway interactions. The New World mammarenavirus Junin virus (JUNV) was reported to induce high levels of pPKR that did not result in increased levels of phosphorylated eukaryotic translation initiation factor 2 alpha (peIF2α), whereas infection with the Old World mammarenavirus LCMV resulted in significantly increased levels of peIF2α despite minimal increased levels of pPKR (37). In addition, infection with the New World mammarenavirus Tacaribe virus (TCRV) resulted in increased levels of both pPKR and peIF2α, which contributed to restricted TCRV multiplication (38). However, other reports documented minimal or negligible activation of both PKR and/or eIF2α in cells infected with mammarenaviruses (39–42). Hence, the need for a detailed characterization of PKR-LCMV interactions is warranted, which is the focus of the present work.
Here, we investigated the interaction of LCMV with the PKR pathway and showed that LCMV infection resulted in minimal increased levels of pPKR, but LCMV multiplication was enhanced in PKR knockout (PKR-KO) cells. However, unexpectedly, pharmacological inhibition of PKR activation resulted in severely restricted multiplication of LCMV, as well as LASV and JUNV. In contrast, cells infected with rLCMV/NP(D382A), impaired in its NP ExoN and anti-T1IFN activities (25), exhibited high levels of pPKR and peIF2α. However, levels of dsRNA remained below detectable levels in rLCMV/NP(D382A)-infected cells. Our findings have uncovered a previously unnoticed pro-viral activity of the PKR pathway during mammarenavirus infection, raising the intriguing possibility that PKR activation might be considered as a potential druggable target to combat infections by human pathogenic mammarenaviruses.
RESULTS
Effect of genetic ablation of PKR on multiplication of LCMV
To determine whether genetic ablation of PKR affected LCMV multiplication, we compared the multi-step growth kinetics of LCMV/wild type (WT) and the ExoN mutant rLCMV/NP(D382A) in WT and PKR-KO A549 cells. We reasoned that LCMV/WT has already a high multiplication fitness in cultured cells, and therefore, ablation of an anti-viral host cell factor would be expected to result in only a modest increase in virus peak titers. In contrast, the fitness of an attenuated LCMV mutant, like rLCMV/NP(D382A), may be more dramatically affected by the removal of an anti-viral host cell factor. We observed that LCMV/WT replicated to high titers in WT and KO A549 lines (Fig. 1A), but virus titers were consistently higher in PKR-KO compared to WT A549 cells. Multiplication of rLCMV/NP(D382A) was severely restricted in WT A549 cells, and production of infectious progeny in tissue culture supernatant (TCS) was only detected at 24 hpi and at very low levels (<102 focus-forming units [FFU]/mL), whereas at 48, 72, and 96 hpi, titers in TCS were below detection levels (Fig. 1A). In contrast, rLCMV/NP(D382A) was able to multiply in PKR KO A549 cells, resulting in titers of infectious progeny in TCS of >103 FFU/mL (Fig. 1A). We further confirmed the effect of PKR on LCMV multiplication in cultured cells by determining levels of LCMV NP RNA using reverse transcription quantitative PCR (RT-qPCR) in WT- and PKR-KO A549-infected cells (Fig. 1B). Consistent with the RT-qPCR results, levels of rLCMV/NP(D382A), but not LCMV/WT, small (S) segment RNA (replication), and NP mRNA (transcription) were drastically reduced in WT compared to PKR-KO A549 cells as determined by northern blotting (Fig. 1C). We next asked if the restricted multiplication of rLCMV/NP(D382A) in A549 WT cells could be overcome by increasing the initial virus input. We infected WT and PKR-KO A549 cells with either LCMV/WT or rLCMV/NP(D382A) using different multiplicities of infection (MOIs) and determined the numbers of NP+ cells at 48 hpi (Fig. 1D). Propagation of LCMV/WT was similarly efficient in WT and PKR-KO A549 cells, even at the lowest (0.05) MOI. In contrast, propagation of rLCMV/NP(D328A) in WT, but not PKR-KO, A549 cells was greatly influenced by the MOI used to initiate infection (Fig. 1Di). Infection of WT A549 cells with rLCMV/NP(D382A) at an MOI of 0.05 or 1.35 resulted in low (6%) and high (90%), respectively, percentage of NP+ cells (Fig. 1Dii). These findings suggest that PKR exerts an anti-viral role during LCMV infection that is counteracted by the ExoN and anti-T1IFN activities of the virus NP.
Fig 1.
Multiplication of LCMV/WT and rLCMV/NP(D382A) in A549 cells. (A) WT or PKR-KO A549 cells (biological triplicate) were infected with LCMV/WT (MOI 0.01) or rLCMV/NP(D382A) (MOI 0.05), and at the indicated hpi virus titers in TCS were determined. Values correspond to the mean ± standard deviation (SD). (B) Determination of NP gene expression levels by RT-qPCR. Total cellular RNA from samples in panel A was isolated using TRI Reagent, and NP gene expression was determined by RT-qPCR. GAPDH was used for normalization utilizing the fold change (2−ΔΔCt) calculation and standardized using the 24-hpi value. Normalized data were plotted as mean ± SD (error bars) with a statistically significant set at P < 0.05. (C) Northern blot analysis. RNA samples from panel B were analyzed by Northern blot. Methylene blue (MB) staining was used to confirm similar transfer efficiency for all RNA samples. Statistically significant values: **P < 0.01, ***P < 0.001, ****P < 0.0001. (D) Effect of MOI on multiplication of rLCMV/NP(D382A). WT or PKR-KO A549 cells were seeded at 2 × 104 cells /well in 96-well plates (two biological replicates, three technical replicates) and infected with LCMV WT or rLCMV/NP(D382A) at the indicated MOIs. At 48 hpi, cells were fixed and analyzed by immunofluorescence (IF). (Di) LCMV-infected cells were identified by IF using the rat monoclonal antibody VL4 to NP, followed by an anti-rat antibody conjugated to AlexaFluor 568. Individual images were obtained using Keyence BZ-X710 imaging system using 1-second exposure time with PlanApo_λ 10 × 0.45/4.00-mm objective lens. Files containing the labeled images were transferred to a laptop for processing the data using ImageJ. PowerPoint (v.2019) was used to compile and arrange the individual images. Adobe Illustrator was used to align the panels within the composite. (Dii) Quantification of A549 WT cells infected with rLCMV/NP(D382A). Cell nuclei were stained with 4',6-diamidino-2-phenylindole (DAPI), and the percentage of infected cells was determined based on the ratio (DAPI++ NP+):DAPI+. ImageJ software was used to quantify the signal, and GraphPad Prism software was used to blot the data. (E and F) Effect of PKR complementation on rLCMV/NP(D382A) multiplication in PKR-KO A549 cells. (E) PKR-KO A549 cells were transfected with pCAGGS empty (pC-E) or pCAGGS expressing PKR-P2A-mCherry (pC-PKR) in suspension, then seeded as indicated per plate format, and on the next day infected with rLCMV/NP(D382A) at an MOI of 0.05. At 48 or 72 hpi, cells were fixed for imaging analysis. Numbers (%) of infected cells were normalized to vehicle-treated infected cells. To ensure accurate comparisons, the signals of the virus (NP) were normalized to the pC-E conditions, which exhibited the highest signal intensity. Conversely, PKR was normalized to PKR transfected cells, as determined by the mCherry signal. This normalization approach was adopted due to the absence of the mCherry signal in both the WT and pC-E conditions. Quantification of colocalization of NP and PKR signals as a function of Pearson’s coefficient, Spearman’s rank coefficient, and area overlap. These measures were obtained from running BIOP JaCoP script (43) using ImageJ software. A control test was done using DAPI images, which were duplicated; Look-Up Table (LUT) values of Region of Interest (ROI) were changed to red and green, then merged, and analyzed using BIOP JaCoP to detect colocalization. The correlation showed 100% agreement. Then the images were split. One image was translated by transformation to variable degrees, re-merged, and re-analyzed using the software to assess the correlation. Correlation levels decreased, depending on the degree of transformation. Duplicate samples were overlayed. (F) PKR-KO A549 cells were transfected with pC-E or pC-PKR in suspension, then seeded at 1 × 106 cells/well into an M6 well plate and on the next day were infected with rLCMV/NP(D382A) at an MOI of 0.05. At 72 hpi, single-cell suspensions were prepared using Accutase, and after two washes with fluorescence activated cell sorting (FACS) buffer, cells were fixed with 4% paraformaldehyde (PFA) for 20 min, permeabilized, then probed with primary and secondary antibodies, followed by two washes after each antibody, re-suspended in FACS buffer, and analyzed using ZE5 analyzer (Bio-Rad). Quantification was done using FlowJo v.10.9, then followed by statistical analysis using GraphPad Prism. Two-way analysis of variance with Šidák correction for multiple comparisons was used. ns, not significant.
To assess whether the enhanced multiplication of rLCMV/NP(D382A) in PKR-KO A549 cells was mainly due to the absence of PKR, we transfected PKR-KO A549 cells with pCAGGS PKR-P2A-mCherrry (pC-PKR), a plasmid expressing PKR tagged with mCherry and containing the self-cleaving peptide (P2A) sequence (44) between the PKR and mCherry open reading frames. This allowed us to use mCherry expression as a surrogate of PKR expression in transfected cells. As a control, we used empty pCAGGS plasmid (pC-E). At 24 h post-transfection, cells were infected with rLCMV/NP(D382A), and at 48 and 72 hpi, infected cells were identified by staining with an NP-specific antibody. We observed a reduced level of rLCMV/NP(D382A) infection at 48 and 72 hpi in A549 PKR-KO cells expressing pC-PKR (mCherry+) compared to cells transfected with the control pC-E plasmid (Fig. 1E), which correlated with greatly reduced colocalization of NP (LCMV infection) and mCherry (PKR expression) as determined by Pearson’s coefficient, Spearman’s rank coefficient, and area overlap measures obtained from running BIOP JaCoP script (43) using ImageJ software. We further validated these findings by flow cytometry (Fig. 1F).
We also examined the role of other interferon (IFN)-induced genes including RNase L and MAVS during LCMV infection and found that MAVS, but not RNAse L, also contributed to the restricted multiplication of rLCMV/NP(D3282A) (data not shown).
Effect of LCMV infection on PKR activation
Activation of PKR requires its autophosphorylation upon binding to viral dsRNA and pPKR negatively regulates translation via phosphorylation of eIF2a. We found that rLCMV/NP(D382A), but not LCMV/WT, activated PKR as determined by the detection of increased levels of pPKR (Fig. 2A) and peIF2α (Fig. 2B). We used infection with Sindbis virus (SINV) as a positive control, as SINV infection activates the PKR pathway (31, 45). Quantification of the immunoblots (IBs) showed that levels of pPKR were fourfold higher in rLCMV/NP(D382A) than LCMV/WT-infected cells, both normalized to levels of pPKR detected in mock-infected control cells (Fig. 2Aii). Using total protein staining to normalize the signal in the IBs, we observed a higher level of pPKR and significantly increased levels of peIF2α in cells infected with rLCMV/NP(D382A), but not in cells infected with LCMV/WT (Fig. 2Bi and Bii).
Fig 2.
(A) PKR activation in LCMV-infected cells. (Ai) WT or PKR-KO A549 cells were infected with LCMV WT (MOI 0.01), rLCMV/NP(D382A) (MOI 0.05), or SINV (MOI 3); at 48 hpi, cell lysates were prepared for Western blot using the indicated antibodies. Samples from SINV-infected cells containing eightfold lower amount of protein, compared to the other lysates, were used as a control (one-eighth). Levels of GAPDH were used for the normalization of samples. Membranes were probed with anti-PKR, anti-pPKR, or anti-GAPDH (primary), anti-GP and horseshoe peroxidase (secondary) antibodies. (Aii) Bands of IB membranes were quantified, and the value of the ratio of pPKR/PKR was plotted. (Bi) WT and PKR-KO A549 cells were infected with LCMV WT (MOI 0.01) or rLCMV/NP(D382) (MOI 0.05). At 48 hpi, cell lysates were prepared for Western blot analysis. IB membrane was probed with anti-PKR, anti-pPKR, or anti-phospho-eIF2α (primary) and the corresponding HRP-conjugated secondary antibodies. The arrow refers to the eIF2α band. (Bii) pPKR:total protein signal ratio. IB signals were quantified using Image Lab software (BioRad). Total protein values were used to normalize the values, and normalized results were plotted. (C) Interferon-dependent PKR-induced expression. (Ci) WT and MAVS-KO A549 cells were seeded in a six-well plate at a density of 1 × 106 cells/well and infected with LCMV WT (MOI 0.01) or rLCMV/NP(D382A) (MOI 0.05). At 48 hpi, cell lysates were prepared for Western blot analysis using anti-PKR and LCMV-GP antibodies. (Cii) IB signals were quantified using Image Lab and normalized to the total protein signal.
PKR is an interferon-stimulated gene (ISG), and LCMV with mutations affecting NP’s ExoN activity (e.g., D382A) has been shown to induce much higher levels of T1IFN than LCMV/WT in infected cells, a process mediated by the cytosolic PRR RIG-I and subsequent activation of the MAVS adaptor and downstream activation of IRF3, a transcriptional factor that promotes induction of IFN-β expression (46). To assess whether the observed higher levels of total PKR protein in rLCMV/NP(D382A)-infected A549 cells was driven by the T1IFN pathway, we compared levels of PKR protein in WT and MAVS-KO A549 cells infected with LCMV/WT or rLCMV/NP(D382A) (Fig. 2C). Levels of PKR were higher in rLCMV/NP(D382A) than LCMV/WT-infected WT A549 cells, whereas both LCMV/WT and rLCMV/NP(D382A)-infected MAVS-KO A549 cells exhibited similar levels of total PKR protein. Consistent with previous results (Fig. 1), levels of LCMV GP2, as determined by Western blot, were slightly higher in MAVS-KO compared to WT-infected A549 cells, whereas infection with rLCMV/NP(D382A) resulted in detectable levels of GP2 expression in MAVS-KO but not in WT A549 cells. These findings supported that T1IFN played a role in driving the expression of total PKR protein in rLCMV/NP(D328A)-ected cells.
We next investigated whether the extremely low increased levels of pPKR in LCMV/WT-infected cells reflected LCMV’s ability to either escape sensing by PKR or actively interfere with the activation of PKR. For this, we infected WT A549 cells with LCMV/WT (MOI 0.5) for 24 h and then superinfected them with SINV (MOI 3). At 24 hpi with SINV, we prepared whole-cell lysates and analyzed them by Western blot (Fig. 3). SINV-mediated activation of PKR, as determined by increased levels of pPKR, was not significantly affected by LCMV infection, indicating that LCMV does not actively inhibit the activation of PKR. SINV infection did not affect LCMV multiplication in A549 cells based on levels of GP2 expression determined by Western blot (Fig. 3A and B) and numbers of LCMV NP+ cells determined by immunofluorescence (IF) (Fig. 3C).
Fig 3.
Effect of LCMV infection on PKR activation. (A) WT A549 cells were infected with LCMV/WT (MOI 0.5) for 24 h and then superinfected with SINV (MOI 3). At 24 hpi with SINV, whole-cell lysates were prepared for Western blot analysis using antibodies to PKR, pPKR, or LCMV-GP. (B) The IB signals were quantified using Image Lab and normalized to total protein signal. (C) WT A549 cells were infected with LCMV/WT (MOI 0.5) for 24 h and then superinfected with SINV (MOI 3). At 24 hpi with SINV, cells were fixed for IF analysis using antibodies against LCMV-NP. Images were taken at ×20 magnification with Keyence BZ-X710.
Role of dsRNA on PKR activation in rLCMV/NP(D382A)-infected cells
We next investigated whether increased levels of dsRNA correlated with increased levels of pPKR observed in rLCMV/NP(D382A) compared to LCMV/WT-infected cells. For this, we infected A549 WT cells with LCMV/WT or rLCMV/NP(D382A) for 24 h, or SINV as a positive control, and used the anti-dsRNA 9D5 antibody to detect dsRNA by IF. Both LCMV/WT- and rLCMV/NP(D382A)-infected cells, as well as mock-infected control cells, had undetectable levels of dsRNA, whereas dsRNA was readily detected in SINV-infected cells (Fig. 4A). We obtained similar results in IFN-deficient Vero E6 cells (Fig. 4B). Levels of dsRNA remained undetectable in Vero E6 cells that had been persistently infected with LCMV for 15 days, but SINV infection of LCMV persistently infected cells resulted in high levels of dsRNA detected by IF (Fig. 4B). Unexpectedly, in contrast to published findings (40), we did not detect dsRNA in cells infected with the Candid#1 live-attenuated strain of the New World Junin virus (JUNV). We obtained similar results using the anti-dsRNA J2 antibody (data not shown). These findings indicated that under our experimental conditions, infection with either WT or NP(D382A) mutant LCMV does not result in levels of dsRNA that can be detected with the validated 9D5 antibody to dsRNA (47, 48).
Fig 4.
Detection of dsRNA. (A) A549 cells were seeded at 1.5 × 104 cells/well in a 96-well plate and were infected with LCMV/WT or rLCMV/NP(D382A) at an MOI of 0.5. At 24 hpi, cells were fixed and analyzed by IF using 9D5 anti-dsRNA antibody. Infection with SINV (MOI 3) was used as a positive control. (B) IFN-deficient Vero E6 cells were infected with LCMV/WT (MOI 0.01), rLCMV/NP(D382A) (MOI 0.05), Candid#1 live-attenuated vaccine strain of JUNV (MOI 0.5), and SINV (MOI 3). At 24 hpi cells were fixed and analyzed by IF as in panel A. Images were taken at ×20 magnification and zoomed in digitally twice. (C) Effect of LCMV persistence on the production of dsRNA. Vero E6 (left) or BHK21 (right) cells were infected with LCMV WT at an MOI of 0.01 in a six-well plate, then expanded to T25 and passaged every 3–4 days along parenteral mock-infected cells. On day 15, cells were seeded into a 96-well plate at a density of 2 × 104 cells/well and incubated for 24 h, then superinfected with SINV (MOI 3). At 24 hpi with SINV, cells were fixed and processed for IF as in panel A. Images were taken using Keyence BZ-X710.
Assessing the contribution of T1IFN to PKR activation in rLCMV/NP(D382A)-infected cells
We have shown that rLCMV/NP(D382A), but not LCMV/WT, can trigger a robust T1IFN response (25). The highly enhanced multiplication of rLCMV/NP(D382A) in PKR-KO compared to WT A549 cells could reflect differences in induction of T1IFN expression between WT and PKR-KO A549 cells following infection with rLCMV/NP(D382A). To examine this possibility, we infected WT or PKR-KO A549 cells with LCMV/WT (MOI 0.01) or rLCMV/NP(D382A) (MOI 0.05) and at 72 h pi collected TCS samples that we used to treat IFN-deficient Vero E6 cells for 24 h, followed by their infection (MOI 0.1) with vesicular stomatitis virus (VSV). At 24 h pi with VSV we fixed the cells and stained them with crystal violet to assess VSV induced cytopathic effect (CPE) (Fig. 5A). Vero E6 cells treated with TCS from either WT or PKR-KO A549 cells infected with rLCMV/NP(D382A) were protected against VSV-induced CPE. As expected, Vero E6 cells treated with TCS from WT or PKR-KO A549 cells infected with LCMV/WT were fully susceptible to VSV induced CPE. Results of quantitative analysis of expression of the ISGs MX1 and ISG15 in Vero E6 cells treated with TCS from A549 cells infected with LCMV/WT or rLCMV/NP(D382A) (Fig. 5B) were consistent with the results of the T1IFN bioassay (Fig. 7A). ISG15 and MX1 genes showed a significant upregulation, ~4- and ~500-folds respectively, in Vero E6 cells treated with TCS from A549 cells infected with NP(D382A), but not LCMV/WT, compared to treatment with TCS from mock-infected A549 cells.
Fig 5.
T1IFN response in PKR-KO A549 cells infected with LCMV. (A) WT and PKR-KO A549 cells were seeded at 1 × 105 cells/well in 24 well plate, and infected on the next day with LCMV/WT (MOI 0.01) or rLCMV/NP(D382A) (MOI 0.05). At 24 h pi, TCS were collected and used to treat Vero E6 cells seeded at 1 × 105 cells/well in a 24-well plate. After 24-h exposure to TCS, Vero E6 cells were infected with VSV at an MOI of 0.1; at 24 hpi with VSV, cells were fixed and stained with crystal violet for 10 min and imaged with BioSpot machine. (B) TCS collected from 48 hpi of A549 WT-infected with LCMV/WT or rLCMV/NP(D382A) was used to treat Vero E6 cells seeded at 1 × 105 cells/well in a 24-well plate. After 24-h exposure to TCS, cells were lysed with TRI Reagent and RNA was collected and pretreated with ezDNAse for 5 min at 37°C, then 10 ng/sample (three technical replicates) for RT-qPCR (two steps: reverse transcription [RT] with random hexamers, followed by quantitative PCR with specific primers). Sybr green was used for RT-qPCR.
Effect of pharmacological inhibition of PKR activation on LCMV multiplication
To assess the effect of PKR inhibition on LCMV multiplication, we examined the effect of the PKR inhibitor C16, and its inactive control C22 (49), on LCMV multiplication. Dose-response assays indicated that C16, but not C22, exerted a dose-dependent inhibitory effect on LCMV multiplication (Fig. 6A). C16 exhibited an EC50 = 0.177 µM and CC50 >20 µM, while C22 exhibited an EC50 >20 µM and CC50 >20 µM (Fig. 6B). Treatment (3 µM) with C16, but not with C22, strongly inhibited propagation of LCMV in A540 cells following infection at low (0.01) MOI (Fig. 6C). Treatment with C16 (3 µM) caused >5 logs reduction in peak titers of rLCMV/green fluorescent protein (GFP) in WT A549 cells (Fig. 6D). A much lower (~ 2 logs) reduction in peak titers of rLCMV/GFP was observed in PKR-KO A549 cells treated with C16, which may reflect C16 off-target effects that affected PKR-independent pathways and factors (50).
Fig 6.
Effect of PKR inhibition on LCMV multiplication. (A) C22 and C16 Dose-response. A549 cells were seeded at 2 × 104 cells/well into a 96-well plate, infected with LCMV/WT (MOI of 0.05) and treated with C16 and C22 at the indicated concentrations. At 96 hpi, cells were fixed, and the numbers of infected cells were determined by IF using a rat mAb to NP. Numbers (%) of infected cells were normalized to vehicle control (VC) infected cells. Results show the percentage of infected WT vs PKR-KO A549 cells (four replicates). (B) C16 and C22 EC50, CC50, and selectivity index (SI). (C) Effect of C16 and C22 on LCMV propagation. A549 cells were seeded at 4 × 104 cells/well into a 96-well plate, infected with LCMV (MOI of 0.05), and treated (3 µM) with either CC16, C22, or VC. At 96 hpi, cells were fixed and stained with DAPI. IF analysis was determined using Keyence BZ-X710. Images were taken at ×10 magnification. (D) Effect of C16 and C22 on the production of virus infectious progeny. A549 cells were seeded at 2 × 105 cells/well in an M12-well plate, infected with LCMV/WT (MOI of 0.01), and treated (3 µM) with C16 or C22 compounds, or with VC. At 72 hpi, TCS was collected, and virus titers were determined by focus-forming assay using Vero E6 cells in a 96-well plate format. The repeated measures analysis of variance with mixed effect analysis and the Dunnet correction for multiple comparisons were used. (E) Time of addition assay. A549 cells were seeded into a 96-well plate at a density of 2 × 104/well. Next day, cells were infected with the single-cycle infectious rLCMV∆GPC/ZsG (MOI = 0.5) and treated (3 µM) with C16 or C22 compounds, or with VC, starting 2 h before (−2 h) or after (+2 h) infection. The LCMV cell entry inhibitor F3406 (5 µM) was used as a control. At 48 hpi, ZsG+ cells were assessed using Cytation 5 reader. Values were normalized to vehicle control-treated and vehicle control-infected cells. (F) Budding assay. HEK293T cells were seeded at a density of on poly-l-lysine-coated wells in a M12-well format. Next day, cells were transfected with either pC.LASV-Z-GLuc or pC.LASV-Z-G2A-GLuc (mutant control) or pCAGGS-Empty(pC-E). At 5 h post-transfection, cells were washed three times and fed with fresh medium containing the indicated drugs and concentrations. At 48 h post-transfection, both TCSs were collected and whole-cell lysis (WCL) was prepared. GLuc activity was determined in TCS and WCL using SteadyGlo Luciferase Pierce: Gaussia Luciferase Glow assay kit utilizing Cytation5 reader. Raw data signal was normalized and then plotted using GraphPad Prism software (v.10). *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001.
To gain insights about the mechanism by which the PKR kinase inhibitor C16 disrupted the LCMV life cycle, we used C16, and its negative control C22, in cell-based assays for different steps of the LCMV life cycle. To distinguish between an effect of C16 on a cell entry or post-entry step of LCMV, we conducted a time of addition assay using a single cycle infectious recombinant LCMV expressing GFP (rLCMV∆GPC/GFP) (51, 52). C16 inhibited multiplication of rLCMV∆GPC/GFP, whether applied 2 h before (−2 hpi) or 2 h after (+2 hpi) infection (Fig. 6E). In contrast, regardless of the time of its addition, C22 did not inhibit rLCMV∆GPC/GFP. As predicted addition at −2 hpi, but not at +2 hpi, of the LCMV entry inhibitor F3406 (51, 52), resulted in inhibition of rLCMV∆GPC/GFP. This observation indicated that C16 inhibited a post cell entry step of the LCMV life cycle. To assess the effect of C16 on mammarenavirus budding, a process driven by the viral matrix protein Z (53), we examined the effect of C16 on the budding activity of the LCMV matrix Z protein using a published cell-based assay where levels of Gaussia luciferase (GLuc) activity serve as an accurate surrogate of Z expression levels (54). C16, but not C22, inhibited Z budding activity (Fig. 6F).
Effect of pharmacological inhibition of PKR on multiplication of other mammarenaviruses
To determine whether our findings could be also extended to other mammarenaviruses, we examined whether multiplication of the New World JUNV and Old World LASV mammarenaviruses was also affected by pharmacological inhibition of PKR activation. For this, we infected A549 WT cells with r3JUNV-GFP (Candid#1 strain) (55) and treated infected cells with C16 (3 µM) or C22 (3 µM) compounds. At the indicated hpi, TCSs were collected, and viral titers were determined. Treatment with compound C16, but not with compound C22, caused three log reductions in virus titers, at 48 and 72 hpi, compared to vehicle control-treated cells (Fig. 7A). Similarly, treatment with C16, but not with C22, inhibited propagation of LASV (Fig. 7B). As with LCMV (Fig. 1), genetic ablation of PKR resulted in enhanced propagation of LASV (Fig. 7C).
Fig 7.
Contribution of PKR to JUNV and LASV multiplication in cultured cells. (A) Treatment with compound C16 inhibits multiplication of JUNV. A549 cells were seeded into an M24 well plate (1.5 × 105 cells/well) and next day (18 h), they were infected with r3JUNV-GFP (Candid#1 strain) (55) at an MOI of 0.05. After 90 min of adsorption (0.2 mL/well), the inoculum was removed; cells were washed once; and fresh medium (0.5 mL/well) containing C16 (3 µM) or C22 (3 µM) compounds, or VC, was added to the cells. At the indicated time points, TCS samples were collected and virus titers were determined by focus-forming unit assay (FFUA) on Vero E6 cells. (B) The PKR inhibitor compound C16 exhibits a dose-dependent inhibitory effect on multiplication of LASV. A549 cells were infected with r3LASV (MOI 0.001). After 1 h of viral adsorption, compounds C16 and C22 or RBV were added at a range of concentrations. At 72 h post-infection, cell culture supernatants were collected and Gluc activity was quantified. Data were normalized by assigning the value of 100 to vehicle control (VC)-treated and r3LASV-infected cells. Cells were then fixed in 10% formalin and imaged for GFP expression using an EVOS M5000 imaging system. (C) Genetic ablation of PKR results in enhanced r3LASV propagation. Parental and PKR KO A549 cells were infected with r3LASV-GFP/Gluc (MOI 0.001). At indicated time points post-infection, TCS samples were collected, and cells were fixed. GFP+ cells, corresponding to LASV-infected cells, were visualized by epifluorescence using an EVOS M5000 imaging (Ci). TCS samples were assayed for their Gluc activity (Cii). Two-way analysis of variance with Šidák correction for multiple comparisons was implemented for statistical analysis. *P < 0.05, **P < 0.01, ***P < 0.001, ****P < 0.0001.
DISCUSSION
The PKR pathway is one of the major arms of cell innate immunity response to viral infections. Accordingly, viruses have evolved different strategies to counteract the PKR pathway. In this work, we have documented that LCMV/WT infection does not result in robust PKR activation, as determined by its autophosphorylation. This finding is consistent with those reported by others showing that at early times of infection and up to 48 hpi, LCMV/WT-infected cells showed negligible activation of PKR (37), as well as results documented with other mammarenaviruses (41). Interestingly, LCMV multiplication was increased in PKR-KO cells, reflected in consistently 5- to 10-fold higher virus titers in PKR-KO than in WT A549 cells. It seems unlikely that LCMV actively interferes with PKR activation as A549-LCMV-infected cells superinfected with SINV exhibited robust PKR activation. In contrast to LCMV/WT, infection of A549 cells with rLCMV/NP(D382A) resulted in robust PKR activation, which correlated with a severely restricted multiplication of rLCMV/NP(D382A) that was alleviated in PKR-KO A549 cells. PKR reconstitution studies using transfection of PKR-KO A549 cells with a PKR-expressing plasmid showed that multiplication of rLCMV/NP(D382) was restricted in PKR-KO A549 cells upon reconstitution of PKR expression, further confirming that PKR can exert an anti-LCMV activity. PKR was found to be involved in apoptosis, the formation of stress granules (56, 57), and IFN-induced cellular necrosis (58). However, neither total protein levels nor numbers of apoptotic cells were significantly affected in cells infected with rLCMV/NP(D382A) despite activation of eIF2α, a finding that warrants further studies.
PKR is activated by conformational change and autophosphorylation upon binding of dsRNA generated during viral genome replication. The 9D5 antibody has been reported to detect dsRNA species ≥40 bp (47, 48, 59) with a motif A2N9A3N9A2 (59), which is present in both GPC (802–826) and L (763–787, 1,007–1,031, 2,970–2,994, 3,711–3,735, and 5,974–5,998) LCMV genes. However, our repeated attempts to detect dsRNA in A549 cells infected with LCMV/WT or rLCMV/NP(D382A) using the 9D5 antibody were unsuccessful, whereas we readily detected dsRNA in cells infected with SINV. Unexpectedly and in contrast with published findings (40), the 9D5 antibody was also unable to detect dsRNA in cells infected with JUNV. Nevertheless, our findings are consistent with earlier observations documenting the detection of dsRNA by the 9D5 antibody in cells infected with dsRNA and positive, but not negative, stranded RNA viruses (60). It is possible that virus-specific dsRNA species generated in LCMV-infected cells are <40 bp in length either because of a fast binding of NP to the newly formed RNA (61) or the lack of immediate supercoiling of RNA (62), which might prevent the formation of the structural binding site for the 9D5 antibody. Alternatively, it might be possible that the 9D5 antibody is targeting a specific domain that is common in dsRNA and +ssRNA supercoiled structures with sizes ≥40 bp (62). The 9D5 antibody has been shown to have an affinity for the peptide motif SIGNAYSMFYDG (63), not present in LCMV proteins, and therefore it cannot be ruled out that the 9D5 antibody binds to a protein target expressed only under specific situations that were not recreated under our experimental conditions.
Our results with PKR-deficient cells supported an expected anti-viral activity of PKR, but intriguingly pharmacological inhibition of PKR activation using compound C16 uncovered a pro-viral activity of PKR. C16, aka imoxin, is a specific PKR inhibitor that has been validated in biochemical (49) and cell-based (64) assays, as well as different animal models of disease including type two diabetes (65), cancer (66), neurodegeneration (67), hypertension (68), rheumatoid arthritis (69), and neuroinflammation (70). C16 has been shown to affect also PKR-independent biochemical intracellular transduction mechanisms (71), and the activity of cyclin-dependent kinases CDK2 and CDK5 (72), which could have contributed to the C16 anti-viral activity against LCMV, LASV, and JUNV. However, we have documented that siRNA-mediated knock-down of the CDK2 activator cyclin A2 protein resulted only in 50% reduction in levels of LCMV multiplication (73), which cannot account for the over five log reductions in production of LCMV infectious progeny caused by C16 treatment we have document in the present work. C16 has been extensively characterized for its inhibitory effect on autophosphorylation of PKR (49). In addition, comprehensive kinomics studies showed that C16 has high specificity for PKR with a Gini score of 0.44 PKR (74). Moreover, other proposed kinase targets of C16, including JNK, MKKs, and GSK3β, are downstream PKR pathway, which may reflect an upstream targeting pf PKR (75–77). Importantly, the C16 closely related compound C22 does not inhibit PKR autophosphorylation (49), which was associated with its lack of anti-mammarenavirus activity, providing additional evidence supporting inhibition of autophosphorylation of PKR as the main factor contributing to the C16 anti-mammarenaviral activity.
PKR may exert functions in LCMV-infected cells that do not require its activation. Thus, PKR has been shown to activate IKK by protein-protein interactions, leading to NF-kB activation and the induction of the T1IFN pathway (78). PKR has also been implicated in the inhibition of gelsolin-mediated regulation of actin dynamics and cytoskeletal cellular functions contributing to innate immunity including restricted virus cell entry, activities that do not appear to require PKR activation via autophosphorylation (79). The actin filament severing activity of gelsolin is inhibited by the oligomeric molecular chaperone CCT (80), and inhibition of the CCT activity resulted in restricted multiplication of LCMV (81), suggesting a connection between PKR, the host cell cytoskeletal activity, and mammarenavirus multiplication.
Similar to findings reported for infectious pancreatic necrosis virus (82) and SARS-CoV-2 (83) treatment with the PKR inhibitor C16 resulted in strong inhibition of LCMV multiplication in WT A549, supporting the role of PKR in LCMV multiplication and uncovering PKR activation as a druggable target for the development of anti-viral drugs against human pathogenic mammarenaviruses, including LCMV, for which evidence indicates it is a neglected human pathogen of clinical significance (4–8) and a serious risk to immunocompromised individuals (9, 10). C16 has shown tolerability and efficacy in different animal models of disease (65–70), supporting its future evaluation, beyond the scope of the present work, in mouse models of LCMV infection.
MATERIALS AND METHODS
Antibodies and compounds
We used the following antibodies at the indicated dilutions: rat mAb VL4 to LCMV NP (BE0106, Bio X cells) 1 in 1,000; mouse mAb 9D5 anti-dsRNA (3361, EMD Millipore) 1 in 2; mouse mAb rJ2 (MABE1134, Millipore/Sigma) 1 in 125; rabbit mAb anti-PKR (D7F7) (12297S, Cell Signaling Technology) 1 in 1,000; antibody (E120) anti-pPKR (phosphor T446) (ab32036, Abcam) 1 in 1,000; rabbit mAb anti-peIF2α (Ser51) (D9G8) XP (3398T, Cell Signaling); Phospho-eIF2α-S51 mouse mAb (AP0692, Abclonal) 1:1,000; mouse Anti-LCMV-GP hybridoma G204 1 in 500, goat anti-rabbit IgG (H + L) cross-adsorbed secondary antibody, Alexa Fluor 488 (A-11008, Thermo Fisher Scientific); goat anti-mouse secondary antibody, Alexa Fluor 568 (Thermo Fisher Scientific); PKR inhibitor C16 (MGF# 15323–1, Thermo Fisher Scientific), and PKR inhibitor negative control (C22) (MGF# 527455–10MG, Fisher Scientific).
Plasmids
pSB819-PKR-hum was a gift from Harmit Malik (Addgene plasmid # 20030; https://www.addgene.org/20030/; RRID: Addgene_20030) (84), and it was used to subclone the human PKR isoform 2 into pCAGGS plasmid preceded by mCherry and separated by P2A self-cleaving peptide sequence (44). The PKR was HA tagged at its C-terminus. HD In-Fusion kit (Takara 638909 In-Fusion HD Cloning Plus) was used for cloning.
Cell lines
A549 (Homo sapiens) [American Type Culture Collection (ATCC), CCL-185], A549 PKR-KO, A549 RNase L-knockout, A549 MAVS-knockout (85), BHK-21 (Mesocricetus auratus) (ATCC, CCL-10), and Vero E6 (Chlorocebus aethiops) (ATCC CRL-1586) cell lines were maintained in Dulbecco’s modified Eagle’s medium (Thermo Fisher Scientific) supplemented with 10% heat-inactivated fetal bovine serum (FBS), 2-mM L-glutamine, 100 µg/mL of streptomycin, and 100 U/mL of penicillin. BHK-21 medium also contained 5% tryptose phosphate broth.
Viruses
The recombinant LCMV, clone 13 (Cl-13), has been described (86); rLCMV/NP(D382A) was rescued via reverse genetics as described (86). Mutation D382A in NP was generated with the In-Fusion cloning method (Takara 638909 In-Fusion HD Cloning Plus). rLCMV/NP(D382A) was grown in a BHK-21 cells, and its identity was confirmed by sequencing. rSINV-GFP was obtained from Matthew Daugherty (UCSD). The recombinant trisegmented (r3) LASV, strain Josiah, expressing GFP and Gluc was rescued as described for other r3MaAv, including LCMV (87), JUNV (88), and TCRV (89), using described LASV reverse genetic approaches (90, 91). Briefly, the mPolI-S LASV plasmid was modified to contain two restrictions sites (BsmBI and BbsI) in opposite orientation (mPolI-S LASV backbone). To generate the mPolI-S1 LASV, the open reading frame (ORF) of EGFP was amplified by PCR and cloned into the BsmBI restriction site of the mPolI-S LASV backbone. Then, the ORF of LASV GPC was amplified by PCR and cloned using BbsI. To generate the mPolI-S2 LASV, the ORF of LASV NP was amplified and cloned into the BsmBI restriction site of the mPolI-S LASV backbone. Then, the ORF of Gluc was amplified and cloned using BbsI. All the plasmids were entirely sequenced by Plasmidsaurus to ensure the absence of unwanted mutations. To generate r3LASV, BHK21 cells in six-well plates were transfected with mPolI-L LASV, mPol-S1 LASV, and mPol-S2 LASV, using LPF2000. At 48 h post-transfection, BHK21 cells were scaled up into a T75 flask, and 48 h after the presence of the r3LASV was verified by GFP expression. Cell culture supernatants were collected to confirm Gluc expression and to generate viral stocks. Stocks of r3LASV were generated and titrated in Vero E6 cells (90, 91). All the experiments with r3LASV were conducted in the biosafety level 4 laboratory at Texas Biomedical Research Institute.
Virus titration
Virus titers were determined by focus-forming assay using Vero E6 cells. Briefly, cells were seeded in 96-well plates at a density of 2 × 104 cells/well and fixed at 20 hpi using 4% paraformaldehyde in phosphate-buffered saline (PBS) for 25 min, washed twice with PBS, permeabilized with 0.3% Triton X-100 containing 3% bovine serum albumin (BSA) in PBS, incubated with the rat mAb VL4 against LCMV NP (Bio X Cell) for 1 h at room temperature, and then incubated with an anti-rat IgG conjugated to Alexa Fluor 568 for 1 h; cells were washed 3× with PBS after each antibody probing step prior to imaging.
RT-qPCR
Cells were infected with LCMV (MOI 0.01) or rLCMV/NP(D382A) (MOI 0.05). Total cellular RNA was isolated using TRI reagent (TR 118) (MRC), and 1 µg was reverse-transcribed to cDNA using the SuperScript IV first-strand synthesis system (Thermo Fisher Scientific). Powerup SYBR (A25742, Life Technologies) was used to amplify LCMV NP and the housekeeping gene GAPDH using the following primers: NP forward (F): 5′ CAGAAATGTTGATGCTGGACTGC-3′ and NP reverse (R): 5′-CAGACCTTGGCTTGCTTTACACAG-3′ (92); GAPDH F: 5′-CATGAGAAGTATGACAACAGCC-3′ and GAPDH R: 5′-TGAGTCCTTCCACGATACC-3′; ISG15 F: 5′-CAGGACGACCTGTTCTGGC-3′ and ISG15 R: 5′-GATTCATGAACACGGTGCTCAGG-3′; and MX1 F: 5′- GCAGCTTCAGAAGGCCATGC-3′ and MX1 R: 5′-CCTTCAGGAACTTCCGCTTGTC-3′.
Western blotting
Cell monolayers were washed with ice-cold PBS, and whole-cell lysates were prepared in cytoplasmic lysis buffer (50-mM Tris HCl, 150-mM NaCl, 1% NP-40, 10% glycerol, and 2-mM EDTA) supplemented with Halt Protease and Phosphatase Inhibitor Cocktails (PI78442, Thermo Fisher Scientific). Lysates were clarified by centrifugation at 10,000 relative centrifugal force (RCF) for 20 min. Samples (24 µg) were denatured by heating for 5 min at 95°C and separated by Stain-Free SDS-PAGE gel (4568096, Bio-Rad). The gel was activated using 1-min setting (Bio-Rad Imager), imaged, then transferred to the low fluorescence polyvinylidene difluoride membrane (1704274, Bio-Rad) that was probed with mAbs to PKR (Cell Signaling Technology) or phospho-PKR-446 (AP1134, Abclonal). Following the incubation with a secondary HRP antibody, the bands were visualized with the chemiluminescent substrate (Thermo Fisher Scientific). The bands’ signals were quantified and analyzed using Image Lab V6 (Bio-Rad).
Northern blotting
RNA samples (5 µg) were fractionated by 2.2-M formaldehyde-agarose (1.2%) gel electrophoresis. The gel was washed once with warm each H2O and 10-mM NaPO4, and RNA transferred in 20× saline sodium citrate (SSC) (3-M sodium chloride, 0.3-M sodium citrate) to a Magnagraph membrane (NJTHYA0010, Osmonics MagnaGraph nylon) using the rapid downward transfer system (TurboBlotter). Membrane-bound RNA was cross-linked by exposure to UV light; the membrane was washed with MilliQ water and stained with methylene blue (MB) to reveal the 18S and 28S RNA plus RNA ladder. After photo documentation, 1% SDS solution was used to remove MB staining, and the membrane was hybridized using QuickHyb (#201220–12 Agilent) to a 32P-labeled dsDNA NP. Hybridization was performed at 65°C overnight. The DNA probe was prepared according to the supplier’s protocol using a DecaPrime kit (Ambion). After overnight hybridization, the membrane was washed twice with 2× SSC–0.2% SDS at 65°C, followed by two washes with 0.2 × SSC–0.2% SDS at 65°C, and then exposed to an X-ray film.
Immunofluorescence
Individual images were obtained using the Keyence BZ-X710. Files containing the labeled images were transferred to a laptop for processing the data using ImageJ. PowerPoint (v.2019) was used to compile and arrange the individual images. Each image was imported individually and organized within the corresponding composite, adjusting the canvas size to ensure a cohesive layout. Adobe Illustrator was used to align the panels within the composite.
Dose-response and cell viability assay
CellTiter 96 AQueous One Solution (G3580, Promega) was used, according to manufacturer protocol, to quantify viable cells. DAPI and GFP signals were quantified after fixation of cells with 4% PFA using a Biotek H4 plate reader, and the Celigo Image Cytometer (Nexcelom). Data were normalized using vehicle-treated infected and mock-infected cells as the highest and lowest signals, respectively. Ribavirin (100 µM) was used as a control of a validated inhibitor of LCMV multiplication. Four replicates were used to quantify each sample.
Flow cytometry
Cells were washed and treated with Accutase (490007–741, VWR), transferred into 15-mL tubes, washed twice with FACS buffer (1× PBS, 2% FBS, 1-mM EDTA), then fixed with 4% PFA for 20 min, and transferred to a 96-well plate (round bottom). After permeabilization for 30 min (Permeabilization Buffer, 00–8333-56, eBioscience), cells were washed and reacted with the indicated antibodies. Cells were washed twice with permeabilization buffer followed by centrifugation for 3 min after each antibody probe. The primary antibody was probed for 1 h and the secondary antibody for 30 min. Cells were re-suspended in FACS buffer for analysis.
Dose-response inhibition of r3LASV
A549 cells (96-welll plate format, quadruplicate) were infected with r3LASV (MOI 0.001). After 1 h of viral adsorption, indicated concentrations of inhibitors were added. At 72 h post-infection, cell culture supernatants were collected, and Gluc activity was quantified according to the manufacturer protocol using Pierce Gaussia Luciferase Glow Assay Kit (16160, Thermo Fisher Scientific) and GloMax plate reader (Promega). Data were normalized to mock-treated, r3LASV-infected cells. Cells were then fixed in 10% formalin and imaged for GFP expression using an EVOS M5000 imaging system (Thermo Fisher Scientific).
Infection of parental and PKR KO A549 cells with r3LASV-GFP
Parental and PKR KO A549 cells (96-welll plate format, triplicates) were infected with r3LASV-GFP/Gluc (MOI 0.001). At indicated time points post-infection, cell culture supernatants were collected and Gluc activity was quantified according to the manufacturer protocol using Pierce Gaussia Luciferase Glow Assay Kit (16160, Thermo Fisher Scientific) and GloMax plate reader (Promega). Data were normalized by subtracting Gluc signal from mock-infected cells. Cells were then fixed in 10% formalin and imaged for GFP using an EVOS M5000 imaging system.
Statistical analysis
Differences that are statistically significant in virus growth and NP gene expression were determined using an unpaired t-test with Welch correction. Bands’ signals of IB were quantified using Image Lab (Bio-Rad). All statistical analyses were conducted using GraphPad Prism software v.9.5.1 (GraphPad). Flow cytometry analyses were performed using FlowJo v.10.9 software.
ACKNOWLEDGMENTS
We thank Olena Shtanko for her help and assistance with the Lassa virus experiments at biosafety level 4.
This research was supported by the National Institutes of Health (NIH)/National Institute of Allergy and Infectious Diseases (grant RO1 AI142985 to J.C.T. and NIH T32 AI007354 to A.S.K.). A.S.K. was supported in part by Open Philanthropy and the Life Sciences Research Foundation.
This is manuscript 30241 from The Scripps Research Institute.
Contributor Information
Juan Carlos de la Torre, Email: juanct@scripps.edu.
Rebecca Ellis Dutch, University of Kentucky College of Medicine, Lexington, Kentucky, USA.
ETHICS APPROVAL
Protocols were approved by Texas Biomedical Research Institute Biosafety and Recombinant DNA committees (BSC21-013 and RDC21-013, respectively).
REFERENCES
- 1. Radoshitzky SR, Buchmeier MJ, de la Torre JC. 2022. Arenaviridae. In Knipe DM, Howley PM, Whelan S (ed), Emerging viruses, 7th ed [Google Scholar]
- 2. Yun NE, Walker DH. 2012. Pathogenesis of Lassa fever. Viruses 4:2031–2048. doi: 10.3390/v4102031 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 3. Grant A, Seregin A, Huang C, Kolokoltsova O, Brasier A, Peters C, Paessler S. 2012. Junín virus pathogenesis and virus replication. Viruses 4:2317–2339. doi: 10.3390/v4102317 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4. Kinori M, Schwartzstein H, Zeid JL, Kurup SP, Mets MB. 2018. Congenital lymphocytic choriomeningitis virus-an underdiagnosed fetal teratogen. J AAPOS 22:79–81. doi: 10.1016/j.jaapos.2017.08.011 [DOI] [PubMed] [Google Scholar]
- 5. Wright R, Johnson D, Neumann M, Ksiazek TG, Rollin P, Keech RV, Bonthius DJ, Hitchon P, Grose CF, Bell WE, Bale JF. 1997. Congenital lymphocytic choriomeningitis virus syndrome: a disease that mimics congenital toxoplasmosis or Cytomegalovirus infection. Pediatrics 100:E9. doi: 10.1542/peds.100.1.e9 [DOI] [PubMed] [Google Scholar]
- 6. Yu JT, Culican SM, Tychsen L. 2006. Aicardi-like chorioretinitis and maldevelopment of the corpus callosum in congenital lymphocytic choriomeningitis virus. J AAPOS 10:58–60. doi: 10.1016/j.jaapos.2005.09.009 [DOI] [PubMed] [Google Scholar]
- 7. Charrel RN, Retornaz K, Emonet S, Noel G, Chaumoitre K, Minodier P, Girard N, Garnier J-M, de Lamballerie X. 2006. Acquired hydrocephalus caused by a variant lymphocytic choriomeningitis virus. Arch Intern Med 166:2044–2046. doi: 10.1001/archinte.166.18.2044 [DOI] [PubMed] [Google Scholar]
- 8. Larsen PD, Chartrand SA, Tomashek KM, Hauser LG, Ksiazek TG. 1993. Hydrocephalus complicating lymphocytic choriomeningitis virus infection. Pediatr Infect Dis J 12:528–531. doi: 10.1097/00006454-199306000-00013 [DOI] [PubMed] [Google Scholar]
- 9. Fischer SA, Graham MB, Kuehnert MJ, Kotton CN, Srinivasan A, Marty FM, Comer JA, Guarner J, Paddock CD, DeMeo DL, et al. 2006. Transmission of lymphocytic choriomeningitis virus by organ transplantation. N Engl J Med 354:2235–2249. doi: 10.1056/NEJMoa053240 [DOI] [PubMed] [Google Scholar]
- 10. Yadav K, Mathur G, Ford B, Miller R. 2014. A case cluster of lymphocytic choriomeningitis virus transmitted via organ transplantation.: abstract# D2381. Transplantation 98:768. doi: 10.1097/00007890-201407151-02624 [DOI] [Google Scholar]
- 11. Salam AP, Cheng V, Edwards T, Olliaro P, Sterne J, Horby P. 2021. Time to reconsider the role of ribavirin in Lassa fever. PLoS Negl Trop Dis 15:e0009522. doi: 10.1371/journal.pntd.0009522 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12. Jensen S, Thomsen AR. 2012. Sensing of RNA viruses: a review of innate immune receptors involved in recognizing RNA virus invasion. J Virol 86:2900–2910. doi: 10.1128/JVI.05738-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13. Lee H-C, Chathuranga K, Lee J-S. 2019. Intracellular sensing of viral genomes and viral evasion. Exp Mol Med 51:1–13. doi: 10.1038/s12276-019-0299-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14. Onomoto K, Onoguchi K, Yoneyama M. 2021. Regulation of RIG-I-like receptor-mediated signaling: interaction between host and viral factors. Cell Mol Immunol 18:539–555. doi: 10.1038/s41423-020-00602-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15. Wong M-T, Chen S-L. 2016. Emerging roles of interferon-stimulated genes in the innate immune response to hepatitis C virus infection. Cell Mol Immunol 13:11–35. doi: 10.1038/cmi.2014.127 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16. Bodur C, Kazyken D, Huang K, Ekim Ustunel B, Siroky KA, Tooley AS, Gonzalez IE, Foley DH, Acosta-Jaquez HA, Barnes TM, Steinl GK, Cho K-W, Lumeng CN, Riddle SM, Myers MG, Fingar DC. 2018. The IKK-related kinase TBK1 activates mTORC1 directly in response to growth factors and innate immune agonists. EMBO J 37:19–38. doi: 10.15252/embj.201696164 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17. Shin CH, Choi D-S. 2019. Essential roles for the non-canonical IκB kinases in linking inflammation to cancer, obesity, and diabetes. Cells 8:178. doi: 10.3390/cells8020178 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18. Falvo JV, Parekh BS, Lin CH, Fraenkel E, Maniatis T. 2000. Assembly of a functional beta interferon enhanceosome is dependent on ATF-2–c-jun heterodimer orientation. Mol Cell Biol 20:4814–4825. doi: 10.1128/MCB.20.13.4814-4825.2000 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19. Nan Y, Wu C, Zhang Y-J. 2017. Interplay between janus kinase/signal transducer and activator of transcription signaling activated by type I interferons and viral antagonism. Front Immunol 8:1758. doi: 10.3389/fimmu.2017.01758 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20. Martínez-Sobrido L, Zúñiga EI, Rosario D, García-Sastre A, de la Torre JC. 2006. Inhibition of the type I interferon response by the nucleoprotein of the prototypic arenavirus lymphocytic choriomeningitis virus. J Virol 80:9192–9199. doi: 10.1128/JVI.00555-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21. García-Sastre A. 2017. Ten strategies of interferon evasion by viruses. Cell Host Microbe 22:176–184. doi: 10.1016/j.chom.2017.07.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22. Ortiz-Riaño E, Cheng BYH, de la Torre JC, Martínez-Sobrido L. 2011. The C-terminal region of lymphocytic choriomeningitis virus nucleoprotein contains distinct and segregable functional domains involved in NP-Z interaction and counteraction of the type I interferon response. J Virol 85:13038–13048. doi: 10.1128/JVI.05834-11 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23. Tsuchida T, Kawai T, Akira S. 2009. Inhibition of IRF3-dependent antiviral responses by cellular and viral proteins. Cell Res 19:3–4. doi: 10.1038/cr.2009.1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24. Chiang H-S, Liu HM. 2018. The molecular basis of viral inhibition of IRF- and STAT-dependent immune responses. Front Immunol 9:3086. doi: 10.3389/fimmu.2018.03086 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25. Martínez-Sobrido L, Emonet S, Giannakas P, Cubitt B, García-Sastre A, de la Torre JC. 2009. Identification of amino acid residues critical for the anti-interferon activity of the nucleoprotein of the prototypic arenavirus lymphocytic choriomeningitis virus. J Virol 83:11330–11340. doi: 10.1128/JVI.00763-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26. Hastie KM, Kimberlin CR, Zandonatti MA, MacRae IJ, Saphire EO. 2011. Structure of the Lassa virus nucleoprotein reveals a dsRNA-specific 3' to 5' exonuclease activity essential for immune suppression. Proc Natl Acad Sci U S A 108:2396–2401. doi: 10.1073/pnas.1016404108 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27. Reynard S, Russier M, Fizet A, Carnec X, Baize S. 2014. Exonuclease domain of the Lassa virus nucleoprotein is critical to avoid RIG-I signaling and to inhibit the innate immune response. J Virol 88:13923–13927. doi: 10.1128/JVI.01923-14 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28. Langevin C, Boudinot P, Collet B. 2019. IFN signaling in inflammation and viral infections: new insights from fish models. Viruses 11:1–17. doi: 10.3390/v11030302 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29. Meurs E, Chong K, Galabru J, Thomas NS, Kerr IM, Williams BR, Hovanessian AG. 1990. Molecular cloning and characterization of the human double-stranded RNA-activated protein kinase induced by interferon. Cell 62:379–390. doi: 10.1016/0092-8674(90)90374-n [DOI] [PubMed] [Google Scholar]
- 30. Li Y, Renner DM, Comar CE, Whelan JN, Reyes HM, Cardenas-Diaz FL, Truitt R, Tan LH, Dong B, Alysandratos KD, Huang J, Palmer JN, Adappa ND, Kohanski MA, Kotton DN, Silverman RH, Yang W, Morrisey EE, Cohen NA, Weiss SR. 2021. SARS-CoV-2 induces double-stranded RNA-mediated innate immune responses in respiratory epithelial-derived cells and cardiomyocytes. Proc Natl Acad Sci U S A 118:e2022643118. doi: 10.1073/pnas.2022643118 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31. Gorchakov R, Frolova E, Williams BRG, Rice CM, Frolov I. 2004. PKR-dependent and -independent mechanisms are involved in translational shutoff during Sindbis virus infection. J Virol 78:8455–8467. doi: 10.1128/JVI.78.16.8455-8467.2004 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32. Ryman KD, Meier KC, Nangle EM, Ragsdale SL, Korneeva NL, Rhoads RE, MacDonald MR, Klimstra WB. 2005. Sindbis virus translation is inhibited by a PKR/RNase L-independent effector induced by alpha/beta interferon priming of dendritic cells. J Virol 79:1487–1499. doi: 10.1128/JVI.79.3.1487-1499.2005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33. Liu Y, Wang M, Cheng A, Yang Q, Wu Y, Jia R, Liu M, Zhu D, Chen S, Zhang S, et al. 2020. The role of host eIF2α in viral infection. Virol J 17:112. doi: 10.1186/s12985-020-01362-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34. Cesaro T, Michiels T. 2021. Inhibition of PKR by viruses. Front Microbiol 12:757238. doi: 10.3389/fmicb.2021.757238 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35. Gao P, Liu Y, Wang H, Chai Y, Weng W, Zhang Y, Zhou L, Ge X, Guo X, Han J, Yang H. 2022. Viral evasion of PKR restriction by reprogramming cellular stress granules. Proc Natl Acad Sci U S A 119:e2201169119. doi: 10.1073/pnas.2201169119 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36. Wang Z, Ren S, Li Q, Royster AD, Lin L, Liu S, Ganaie SS, Qiu J, Mir S, Mir MA. 2021. Hantaviruses use the endogenous host factor P58IPK to combat the PKR antiviral response. PLoS Pathog 17:e1010007. doi: 10.1371/journal.ppat.1010007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37. King BR, Hershkowitz D, Eisenhauer PL, Weir ME, Ziegler CM, Russo J, Bruce EA, Ballif BA, Botten J. 2017. A map of the arenavirus nucleoprotein-host protein interactome reveals that Junín virus selectively impairs the antiviral activity of double-stranded RNA-activated protein kinase (PKR). J Virol 91:e00763-17. doi: 10.1128/JVI.00763-17 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38. Moreno H, Kunz S. 2021. The protein kinase receptor modulates the innate immune response against Tacaribe virus. Viruses 13:1313. doi: 10.3390/v13071313 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39. Huang C, Kolokoltsova OA, Mateer EJ, Koma T, Paessler S. 2017. Highly pathogenic new world arenavirus infection activates the pattern recognition receptor protein kinase R without attenuating virus replication in human cells. J Virol 91:e01090-17. doi: 10.1128/JVI.01090-17 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40. Mateer EJ, Maruyama J, Card GE, Paessler S, Huang C. 2020. Lassa virus, but not highly pathogenic new world arenaviruses, restricts immunostimulatory double-stranded RNA accumulation during infection. J Virol 94:e02006-19. doi: 10.1128/JVI.02006-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41. Kainulainen MH, Spengler JR, Welch SR, Coleman-McCray JD, Harmon JR, Scholte FEM, Goldsmith CS, Nichol ST, Albariño CG, Spiropoulou CF. 2019. Protection from lethal Lassa disease can be achieved both before and after virus exposure by administration of single-cycle replicating Lassa virus replicon particles. J Infect Dis 220:1281–1289. doi: 10.1093/infdis/jiz284 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42. Brinkmann NM, Hoffmann C, Wurr S, Pallasch E, Hinzmann J, Ostermann E, Brune W, Eskes ME, Jungblut L, Günther S, Unrau L, Oestereich L. 2022. Understanding host-virus interactions: assessment of innate immune responses in Mastomys natalensis cells after arenavirus infection. Viruses 14:1986. doi: 10.3390/v14091986 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43. Bolte S, Cordelières FP. 2006. A guided tour into subcellular colocalization analysis in light microscopy. J Microsc 224:213–232. doi: 10.1111/j.1365-2818.2006.01706.x [DOI] [PubMed] [Google Scholar]
- 44. Szymczak-Workman AL, Vignali KM, Vignali DAA. 2012. Design and construction of 2A peptide-linked multicistronic vectors. Cold Spring Harb Protoc 2012:199–204. doi: 10.1101/pdb.ip067876 [DOI] [PubMed] [Google Scholar]
- 45. Comar CE, Goldstein SA, Li Y, Yount B, Baric RS, Weiss SR. 2019. Antagonism of dsRNA-induced innate immune pathways by NS4a and NS4b accessory proteins during MERS coronavirus infection. mBio 10:e00319-19. doi: 10.1128/mBio.00319-19 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46. Pythoud C, Rothenberger S, Martínez-Sobrido L, de la Torre JC, Kunz S. 2015. Lymphocytic choriomeningitis virus differentially affects the virus-induced type I interferon response and mitochondrial apoptosis mediated by RIG-I/MAVS. J Virol 89:6240–6250. doi: 10.1128/JVI.00610-15 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47. Gao Y, Chen S, Halene S, Tebaldi T. 2021. Transcriptome-wide quantification of double-stranded RNAs in live mouse tissues by dsRIP-Seq. STAR Protoc 2:100366. doi: 10.1016/j.xpro.2021.100366 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48. O’Brien CA, Hobson-Peters J, Yam AWY, Colmant AMG, McLean BJ, Prow NA, Watterson D, Hall-Mendelin S, Warrilow D, Ng M-L, Khromykh AA, Hall RA. 2015. Viral RNA intermediates as targets for detection and discovery of novel and emerging mosquito-borne viruses. PLoS Negl Trop Dis 9:e0003629. doi: 10.1371/journal.pntd.0003629 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49. Jammi NV, Whitby LR, Beal PA. 2003. Small molecule inhibitors of the RNA-dependent protein kinase. Biochem Biophys Res Commun 308:50–57. doi: 10.1016/s0006-291x(03)01318-4 [DOI] [PubMed] [Google Scholar]
- 50. Weintraub S, Yarnitzky T, Kahremany S, Barrera I, Viskind O, Rosenblum K, Niv MY, Gruzman A. 2016. Design and synthesis of novel protein kinase R (PKR) inhibitors. Mol Divers 20:805–819. doi: 10.1007/s11030-016-9689-4 [DOI] [PubMed] [Google Scholar]
- 51. Rodrigo W, de la Torre JC, Martínez-Sobrido L. 2011. Use of single-cycle infectious lymphocytic choriomeningitis virus to study hemorrhagic fever arenaviruses. J Virol 85:1684–1695. doi: 10.1128/JVI.02229-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52. Wan W, Zhu S, Li S, Shang W, Zhang R, Li H, Liu W, Xiao G, Peng K, Zhang L. 2021. High-throughput screening of an FDA-approved drug library identifies inhibitors against arenaviruses and SARS-CoV-2. ACS Infect Dis 7:1409–1422. doi: 10.1021/acsinfecdis.0c00486 [DOI] [PubMed] [Google Scholar]
- 53. Perez M, Craven RC, de la Torre JC. 2003. The small RING finger protein Z drives arenavirus budding: implications for antiviral strategies. Proc Natl Acad Sci U S A 100:12978–12983. doi: 10.1073/pnas.2133782100 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54. Capul AA, de la Torre JC. 2008. A cell-based luciferase assay amenable to high-throughput screening of inhibitors of arenavirus budding. Virology 382:107–114. doi: 10.1016/j.virol.2008.09.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55. Emonet SF, Seregin AV, Yun NE, Poussard AL, Walker AG, de la Torre JC, Paessler S. 2011. Rescue from cloned cDNAs and in vivo characterization of recombinant pathogenic romero and live-attenuated Candid #1 strains of Junin virus, the causative agent of Argentine hemorrhagic fever disease. J Virol 85:1473–1483. doi: 10.1128/JVI.02102-10 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56. Gil J, Esteban M. 2000. The interferon-induced protein kinase (PKR), triggers apoptosis through FADD-mediated activation of caspase 8 in a manner independent of Fas and TNF-alpha receptors. Oncogene 19:3665–3674. doi: 10.1038/sj.onc.1203710 [DOI] [PubMed] [Google Scholar]
- 57. Lindquist ME, Mainou BA, Dermody TS, Crowe JE. 2011. Activation of protein kinase R is required for induction of stress granules by respiratory syncytial virus but dispensable for viral replication. Virology 413:103–110. doi: 10.1016/j.virol.2011.02.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58. Thapa RJ, Nogusa S, Chen P, Maki JL, Lerro A, Andrake M, Rall GF, Degterev A, Balachandran S. 2013. Interferon-induced RIP1/RIP3-mediated necrosis requires PKR and is licensed by FADD and caspases. Proc Natl Acad Sci U S A 110:E3109–E3118. doi: 10.1073/pnas.1301218110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59. Bonin M, Oberstrass J, Lukacs N, Ewert K, Oesterschulze E, Kassing R, Nellen W. 2000. Determination of preferential binding sites for anti-dsRNA antibodies on double-stranded RNA by scanning force microscopy. RNA 6:563–570. doi: 10.1017/s1355838200992318 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60. Weber F, Wagner V, Rasmussen SB, Hartmann R, Paludan SR. 2006. Double-stranded RNA is produced by positive-strand RNA viruses and DNA viruses but not in detectable amounts by negative-strand RNA viruses. J Virol 80:5059–5064. doi: 10.1128/JVI.80.10.5059-5064.2006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61. Liu C-L, Hung H-C, Lo S-C, Chiang C-H, Chen I-J, Hsu JT-A, Hou M-H. 2016. Using mutagenesis to explore conserved residues in the RNA-binding groove of influenza A virus nucleoprotein for antiviral drug development. Sci Rep 6:21662. doi: 10.1038/srep21662 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62. Stagno JR, Ma B, Li J, Altieri AS, Byrd RA, Ji X. 2012. Crystal structure of a plectonemic RNA supercoil. Nat Commun 3:901. doi: 10.1038/ncomms1903 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63. Maccari G, Genoni A, Sansonno S, Toniolo A. 2016. Properties of two enterovirus antibodies that are utilized in diabetes research. Sci Rep 6:24757. doi: 10.1038/srep24757 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64. Gilfoy FD, Mason PW. 2007. West Nile virus-induced interferon production is mediated by the double-stranded RNA-dependent protein kinase PKR. J Virol 81:11148–11158. doi: 10.1128/JVI.00446-07 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65. Nakamura T, Arduini A, Baccaro B, Furuhashi M, Hotamisligil GS. 2014. Small-molecule inhibitors of PKR improve glucose homeostasis in obese diabetic mice. Diabetes 63:526–534. doi: 10.2337/db13-1019 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66. Watanabe T, Imamura T, Hiasa Y. 2018. Roles of protein kinase R in cancer: potential as a therapeutic target. Cancer Sci 109:919–925. doi: 10.1111/cas.13551 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67. Ingrand S, Barrier L, Lafay-Chebassier C, Fauconneau B, Page G, Hugon J. 2007. The oxindole/imidazole derivative C16 reduces in vivo brain PKR activation. FEBS Lett 581:4473–4478. doi: 10.1016/j.febslet.2007.08.022 [DOI] [PubMed] [Google Scholar]
- 68. Kalra J, Dasari D, Bhat A, Mangali S, Goyal SG, Jadhav KB, Dhar A. 2020. PKR inhibitor imoxin prevents hypertension, endothelial dysfunction and cardiac and vascular remodelling in L-NAME-treated rats. Life Sci 262:118436. doi: 10.1016/j.lfs.2020.118436 [DOI] [PubMed] [Google Scholar]
- 69. Wang WJ, Yin SJ, Rong RQ. 2015. PKR and HMGB1 expression and function in rheumatoid arthritis. Genet Mol Res 14:17864–17870. doi: 10.4238/2015.December.22.11 [DOI] [PubMed] [Google Scholar]
- 70. Tronel C, Page G, Bodard S, Chalon S, Antier D. 2014. The specific PKR inhibitor C16 prevents apoptosis and IL-1β production in an acute excitotoxic rat model with a neuroinflammatory component. Neurochem Int 64:73–83. doi: 10.1016/j.neuint.2013.10.012 [DOI] [PubMed] [Google Scholar]
- 71. Couturier J, Paccalin M, Morel M, Terro F, Milin S, Pontcharraud R, Fauconneau B, Page G. 2011. Prevention of the β-amyloid peptide-induced inflammatory process by inhibition of double-stranded RNA-dependent protein kinase in primary murine mixed co-cultures. J Neuroinflammation 8:72. doi: 10.1186/1742-2094-8-72 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Chen H-M, Wang L, D’Mello SR. 2008. A chemical compound commonly used to inhibit PKR, {8-(imidazol-4-ylmethylene)-6H-azolidino[5,4-g]benzothiazol-7-one}, protects neurons by inhibiting cyclin-dependent kinase. Eur J Neurosci 28:2003–2016. doi: 10.1111/j.1460-9568.2008.06491.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73. Kim Y-J, Witwit H, Cubitt B, de la Torre JC. 2021. Inhibitors of anti-apoptotic Bcl-2 family proteins exhibit potent and broad-spectrum anti-mammarenavirus activity via cell cycle arrest at G0/G1 phase. J Virol 95:e0139921. doi: 10.1128/JVI.01399-21 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74. Anastassiadis T, Deacon SW, Devarajan K, Ma H, Peterson JR. 2011. Comprehensive assay of kinase catalytic activity reveals features of kinase inhibitor selectivity. Nat Biotechnol 29:1039–1045. doi: 10.1038/nbt.2017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75. Piazzi M, Bavelloni A, Faenza I, Blalock W. 2020. Glycogen synthase kinase (GSK)-3 and the double-strand RNA-dependent kinase, PKR: when two kinases for the common good turn bad. Biochim Biophys Acta Mol Cell Res 1867:118769. doi: 10.1016/j.bbamcr.2020.118769 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Williams BR. 1999. PKR; a sentinel kinase for cellular stress. Oncogene 18:6112–6120. doi: 10.1038/sj.onc.1203127 [DOI] [PubMed] [Google Scholar]
- 77. García MA, Gil J, Ventoso I, Guerra S, Domingo E, Rivas C, Esteban M. 2006. Impact of protein kinase PKR in cell biology: from antiviral to antiproliferative action. Microbiol Mol Biol Rev 70:1032–1060. doi: 10.1128/MMBR.00027-06 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78. Bonnet MC, Weil R, Dam E, Hovanessian AG, Meurs EF. 2000. PKR stimulates NF-κB irrespective of its kinase function by interacting with the IκB kinase complex. Mol Cell Biol 20:4532–4542. doi: 10.1128/MCB.20.13.4532-4542.2000 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79. Irving AT, Wang D, Vasilevski O, Latchoumanin O, Kozer N, Clayton AHA, Szczepny A, Morimoto H, Xu D, Williams BRG, Sadler AJ. 2012. Regulation of actin dynamics by protein kinase R control of gelsolin enforces basal innate immune defense. Immunity 36:795–806. doi: 10.1016/j.immuni.2012.02.020 [DOI] [PubMed] [Google Scholar]
- 80. Svanström A, Grantham J. 2016. The molecular chaperone CCT modulates the activity of the actin filament severing and capping protein gelsolin in vitro. Cell Stress Chaperones 21:55–62. doi: 10.1007/s12192-015-0637-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81. Sakabe S, Witwit H, Khafaji R, Cubitt B, de la Torre JC. 2023. Chaperonin TRiC/CCT participates in mammarenavirus multiplication in human cells via interaction with the viral nucleoprotein. J Virol 97:e0168822. doi: 10.1128/jvi.01688-22 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82. Gamil AAA, Xu C, Mutoloki S, Evensen Ø. 2016. PKR activation favors infectious pancreatic necrosis virus replication in infected cells. Viruses 8:173. doi: 10.3390/v8060173 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83. Jain S, Rego S, Park S, Liu Y, Parn S, Savsani K, Perlin DS, Dakshanamurthy S. 2022. RNASeq profiling of COVID19-infected patients identified an EIF2AK2 inhibitor as a potent SARS-CoV-2 antiviral. Clin Transl Med 12:e1098. doi: 10.1002/ctm2.1098 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84. Elde NC, Child SJ, Geballe AP, Malik HS. 2009. Protein kinase R reveals an evolutionary model for defeating viral mimicry. Nature 457:485–489. doi: 10.1038/nature07529 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85. Li Y, Banerjee S, Goldstein SA, Dong B, Gaughan C, Rath S, Donovan J, Korennykh A, Silverman RH, Weiss SR. 2017. Ribonuclease L mediates the cell-lethal phenotype of double-stranded RNA editing enzyme ADAR1 deficiency in a human cell line. Elife 6:e25687. doi: 10.7554/eLife.25687 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86. Sánchez AB, de la Torre JC. 2006. Rescue of the prototypic arenavirus LCMV entirely from plasmid. Virology 350:370–380. doi: 10.1016/j.virol.2006.01.012 [DOI] [PubMed] [Google Scholar]
- 87. Cheng BYH, Ortiz-Riaño E, de la Torre JC, Martínez-Sobrido L. 2015. Arenavirus genome rearrangement for the development of live attenuated vaccines. J Virol 89:7373–7384. doi: 10.1128/JVI.00307-15 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88. Cheng BYH, Ortiz-Riaño E, de la Torre JC, Martínez-Sobrido L. 2013. Generation of recombinant arenavirus for vaccine development in FDA-approved Vero cells. J Vis Exp 50662:50662. doi: 10.3791/50662 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89. Ye C, de la Torre JC, Martínez-Sobrido L. 2020. Development of reverse genetics for the prototype new world mammarenavirus Tacaribe virus. J Virol 94:e01014-20. doi: 10.1128/JVI.01014-20 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90. Yun NE, Seregin AV, Walker DH, Popov VL, Walker AG, Smith JN, Miller M, de la Torre JC, Smith JK, Borisevich V, Fair JN, Wauquier N, Grant DS, Bockarie B, Bente D, Paessler S. 2013. Mice lacking functional STAT1 are highly susceptible to lethal infection with Lassa virus. J Virol 87:10908–10911. doi: 10.1128/JVI.01433-13 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91. Martínez-Sobrido L, Paessler S, de la Torre JC. 2017. Lassa virus reverse genetics. Methods Mol Biol 1602:185–204. doi: 10.1007/978-1-4939-6964-7_13 [DOI] [PubMed] [Google Scholar]
- 92. McCausland MM, Crotty S. 2008. Quantitative PCR (QPCR) technique for detecting lymphocytic choriomeningitis virus (LCMV) in vivo. J Virol Methods 147:167–176. doi: 10.1016/j.jviromet.2007.08.025 [DOI] [PMC free article] [PubMed] [Google Scholar]







