Skip to main content
PLOS Neglected Tropical Diseases logoLink to PLOS Neglected Tropical Diseases
. 2022 Jun 27;16(6):e0010559. doi: 10.1371/journal.pntd.0010559

Zika virus infection drives epigenetic modulation of immunity by the histone acetyltransferase CBP of Aedes aegypti

Anderson de Mendonça Amarante 1,2, Isabel Caetano de Abreu da Silva 1,2, Vitor Coutinho Carneiro 3, Amanda Roberta Revoredo Vicentino 1, Marcia de Amorim Pinto 1, Luiza Mendonça Higa 4, Kanhu Charan Moharana 5, Octavio A C Talyuli 1,2, Thiago Motta Venancio 2,5, Pedro Lagerblad de Oliveira 1,2, Marcelo Rosado Fantappié 1,2,*
Editor: Jason L Rasgon6
PMCID: PMC9269902  PMID: 35759510

Abstract

Epigenetic mechanisms are responsible for a wide range of biological phenomena in insects, controlling embryonic development, growth, aging and nutrition. Despite this, the role of epigenetics in shaping insect-pathogen interactions has received little attention. Gene expression in eukaryotes is regulated by histone acetylation/deacetylation, an epigenetic process mediated by histone acetyltransferases (HATs) and histone deacetylases (HDACs). In this study, we explored the role of the Aedes aegypti histone acetyltransferase CBP (AaCBP) after infection with Zika virus (ZIKV), focusing on the two main immune tissues, the midgut and fat body. We showed that the expression and activity of AaCBP could be positively modulated by blood meal and ZIKV infection. Nevertheless, Zika-infected mosquitoes that were silenced for AaCBP revealed a significant reduction in the acetylation of H3K27 (CBP target marker), followed by downmodulation of the expression of immune genes, higher titers of ZIKV and lower survival rates. Importantly, in Zika-infected mosquitoes that were treated with sodium butyrate, a histone deacetylase inhibitor, their capacity to fight virus infection was rescued. Our data point to a direct correlation among histone hyperacetylation by AaCBP, upregulation of antimicrobial peptide genes and increased survival of Zika-infected-A. aegypti.

Author summary

Pathogens have coevolved with mosquitoes to optimize transmission to hosts. As natural vectors, mosquitoes are permissive to and allow systemic and persistent arbovirus infection, which intriguingly does not result in dramatic pathological sequelae that affect their lifespan. In this regard, mosquitoes have evolved mechanisms to tolerate persistent infection and develop efficient antiviral strategies to restrict viral replication to nonpathogenic levels. There is a great deal of evidence supporting the implication of epigenetics in the modulation of the biological interaction between hosts and pathogens. This study reveals that Zika virus infection positively modulates the expression and activity of A. aegypti histone acetyltransferase CBP (AaCBP). This study shows that AaCBP plays a role in the activation of immune-responsive genes to limit Zika virus replication. This first description that Zika virus infection has epigenomic consequences in the regulation of A. aegypti immunity opens a new avenue for research on mosquito factors that can drive vector competence.

Introduction

Mosquitoes are primary vectors of a variety of human pathogens throughout the world. Aedes aegypti mosquitoes can develop long-lasting viral infections and carry high viral loads, which make them efficient vectors for the transmission of arboviruses such as Zika virus (ZIKV) [1].

Host-pathogen interactions are among the most plastic and dynamic systems in nature. In this regard, epigenetic modifications can offer an accessory source of fast-acting, reversible and readily available phenotypic variation that can be directly shaped by both host and pathogen selection pressures [2,3]. One of the hallmarks in the study of host gene regulation is to elucidate how specific sets of genes are selected for expression in response to pathogen infection.

Understanding the interactions between the mosquito immune system and viruses is critical for the development of effective control strategies against these diseases. Mosquitoes have conserved immune pathways that limit infections by viral pathogens [4,5]. Mosquito antiviral defense is regulated by RNA interference (RNAi), Janus kinase/signal transducer (JAK-STAT), Toll, the immune deficiency (IMD) and MAPK immune pathways [46].

The Toll pathway has been shown to play the most important role in controlling ZIKV infections [4]. In this context, gene expression analysis of ZIKV-infected mosquitoes has indicated that Toll pathway-related genes are highly upregulated in Zika infection when compared to other immune pathways [4,5].

Eukaryotic gene expression is controlled by the functions of cis-DNA elements, enhancers and promoters, which are bound by transcription factors, in combination with the organization of the chromatin [7]. Although it is still poorly understood how chromatin-associated processes participate in the regulation of gene transcription in the context of pathogen-vector interactions, it has been previously shown that Plasmodium falciparum infection induces significant chromatin changes in the Anopheles gambiae mosquitoes [3]. This study identified infection-responsive genes showing differential enrichment in various histone modifications at the promoter sites [3].

The transcriptional coactivator CREB-binding protein (CBP), and its paralog p300, play a central role in coordinating and integrating multiple signal-dependent events with the transcription apparatus, allowing the appropriate level of gene activity to occur in response to diverse stimuli [8]. CBP proteins do not specifically interact with the promoter elements of target genes, but they are recruited to promoters by interaction with DNA-bound transcription factors, which directly interact with the RNA polymerase II complex [8,9]. A key property of CBP is the presence of histone acetyltransferase (HAT) activity, which endows the enzyme with the capacity to influence chromatin activity by the acetylation of histones [8].

Consistent with its function as a transcriptional coactivator, CBP plays crucial roles in embryogenesis [10], development [11], differentiation [11,12], oncogenesis [11,13] and immunity [9]. Although most of these studies have been conducted in mammalian systems, nonvector insect models have also made important contributions to the biological functions of CBP [1417].

The present work describes the functional characterization of a histone acetyltransferase in an insect vector and provides indications that Zika virus infection exerts epigenomic consequences in regulating A. aegypti immunity.

Materials and methods

Ethics statement

All animal care and experimental protocols were conducted in accordance with the guidelines of the Committee for Evaluation of Animal Use for Research–CEUA of the Federal University of Rio de Janeiro (UFRJ). The protocols were approved by CEUA-UFRJ under the registration number IBQM149/19. Technicians in the animal facility at the Instituto de Bioquímica Médica Leopoldo de Meis (UFRJ) carried out all protocols related to rabbit husbandry under strict guidelines to ensure careful and consistent animal handling.

Mosquito rearing and cell culture

Aedes aegypti (Liverpool black-eyed strain) were raised in a mosquito rearing facility at the Federal University of Rio de Janeiro, Brazil, under a 12-h light/dark cycle at 28°C and 70–80% relative humidity. Larvae were fed dog chow, and adults were maintained in a cage and given a solution of 10% sucrose ad libitum. Females 7–10 days posteclosion were used in the experiments. When mentioned, mosquitoes were artificially fed heparinized rabbit blood. For virus infection, mosquitoes were fed using water-jacketed artificial feeders maintained at 37°C sealed with parafilm membranes. Alternatively, mosquitoes were infected by intrathoracic injections of 69 nL containing 60 Plaque Forming Units (PFUs) of ZIKV.

Female midguts or fat bodies were dissected 24, 48, 72 or 96 h after feeding for RNA sample preparation.

The A. aegypti embryonic cell line Aag2 was maintained in Schneider medium (Merk) supplemented with 5% FBS (LGC, Brazil) and 1% penicillin/streptomycin/amphotericin B. Aag2 cells were incubated at 28°C.

ZIKV infection and virus titration

ZIKV strain Pernambuco (ZIKV strain ZIKV/H.sapiens/Brazil/PE243/201) [18] was propagated in the A. albobictus C6/36 cell line, and titers were determined by plaque assay on Vero cells. ZIKV was propagated in C6/36 cells for 6–8 days; virus was then harvested, filtered through 0.22 and 0.1μm membranes, respectively and stored at -80°C. Mosquitoes were intrathoracically infected by microinjections [19] of 69 nL of virus, containing 60 PFUs.

Midguts and fat bodies were dissected, individually collected, and stored at– 80°C until use for plaque assays. Virus titration was performed as described previously [20]. Plates were incubated for 4–5 days, fixed and stained with a methanol/acetone and 1% crystal violet mixture, and washed, after which the plaque forming units (PFUs) were counted.

AaCBP gene knockdown by RNAi

Double-stranded RNA (dsRNA) was synthesized from templates amplified from cDNA of adult female mosquitoes using specific primers containing a T7 tail (S1 Table). The in vitro dsRNA transcription reaction was performed following the manufacturer’s instructions (Ambion MEGAscript RNAi). Two different dsRNA products (dsAaCBP1 and dsAaCBP2) were PCR amplified based on the coding sequence of A. aegypti CBP (GenBank accession number XP_011493407.2), using the oligonucleotides listed in S1 Table. The irrelevant control gene luciferase (dsLuc) was amplified from the luciferase T7 control plasmid (Promega). Female mosquitoes were injected intrathoracically [21] with dsRNA (0.4 μg) with a microinjector (NanoJect II Autonanoliter injector, Drummond Scientific, USA). Injected mosquitoes were maintained at 28°C, and 70–80% humidity, with 10% sucrose provided ad libitum.

Nucleic acids isolation and quantitative real-time PCR

Total RNA was isolated from whole bodies, midguts or fat bodies of adult females, using the RiboPure kit (Ambion) followed by DNase treatment (Ambion) and cDNA synthesis (SuperScript III First-Strand Synthesis System, Invitrogen), following the manufacturer’s instructions. Quantitative reverse transcription gene amplifications (qRT-PCR) were performed with StepOnePlus Real-Time PCR System (Applied Biosystems) using the Power SYBR Green PCR Master Mix (Applied Biosystems). The comparative Ct method [22] was used to compare mRNA abundance. In all qRT-PCR analyses, the A. aegypti ribosomal protein 49 gene (Rp49) was used as an endogenous control [23]. All oligonucleotide sequences used in qRT-PCR assays are listed in the S1 Table.

The content of the mosquito intestinal microbiota was analyzed by genomic DNA qPCR. Genomic DNA was isolated using the DNeasy Blood and Tissue Kit (Qiagen), following the manufacturer’s instructions. Amplifications were carried out using specific primers of the ribosomal genes 16S and 18S (S1 Table), for bacteria or fungi, respectively. The A. aegypti ribosomal protein 49 gene (Rp49) was used as an endogenous control. Quantifications were carried out using the comparative Ct method. Fifty nanograms of genomic DNA from a pool of 15 midguts from female mosquitoes was used as a template. For midgut dissection, mosquitoes were sterilized in 20% hypochlorite solution, followed by 70% ethanol for 30 seconds in each solution. Sterilized midguts were washed in sterile 1X PBS solution.

Western blotting

Protein extracts were prepared as previously described [24,25]. Briefly, A. aegypti total protein extracts were carried out by homogenizing adult female mosquitoes or Aag2 cells in TBS containing a protease inhibitor cocktail (Sigma). Proteins were recovered from the supernatant by centrifugation at 14.000xg, for 15 min. at 4°C. Protein concentration was determined by the Bradford Protein Assay (Bio-Rad). Western blots were carried out using secondary antibody (Immunopure goat anti-mouse, #31430). The primary monoclonal antibodies (ChIP grade) used were anti-H3 pan acetylated (Sigma-Aldrich #06–599), anti-H3K9ac (Cell Signaling Technology #9649) and Anti-H3K27ac (Cell Signaling Technology, #8173), according to the manufacture’s instructions. For all antibodies, a 1:1000 dilution was used. For normalization of the signals across the samples, an anti-histone H3 antibody (Cell Signaling Technology, #14269) was used.

Statistical analysis

Survival curves were generated with the Kaplam-meier analysis using the log-rank test. All other analyses were performed with the GraphPad Prism statistical software package (Prism version 6.0, GraphPad Software, Inc., La Jolla, CA). Asterisks indicate significant differences (*, p < 0.05; **, p < 0.01; ***, p < 0.001; ns, not significant).

Genome-wide identification of lysine acetyltransferases in A. aegypti

We obtained the latest A. aegypti proteome and functional annotations (AaegL5.0) from NCBI RefSeq (https://ftp.ncbi.nlm.nih.gov/genomes/all/GCF/002/204/515/GCF_002204515.2_AaegL5.0). The amino acid sequences of the 23 D. melanogaster lysine acetyltransferases (DmKATs) reported by Feller et al 2015 were obtained from FlyBase. The KAT orthologs in A. aegypti were identified using BlastP (e-value ≤1e-10, identity ≥25% and query coverage ≥50%) [26]. We further validated the presence of various acetyltransferase domains in these KATs. Conserved domains were predicted in putative KATs using hmmer (e-value ≤ 0.01; http://hmmer.org/) and the PFAM-A database. Following the nomenclature of DmKATs and the conserved domain architectures, AeKATs were classified into 5 major sub-families: HAT1, Tip60, MOF, HBO1, GCN5 and CBP [27]. Conserved domain architectures were rendered with DOG (Domain Graph, version 1.0) [28]. Multiple sequence alignment of AeKATs was performed using using Clustalw [29].

Results

Aedes aegypti has an ortholog of the CREB-binding protein (CBP) and five additional putative histone acetyltransferases

The A. aegypti transcriptional coactivator CBP (AaCBP) contains all five canonical structural and functional domains (Fig 1) of the CBP family [30]. The TAZ, KIX and CREB are the protein interaction domains that mediate interactions with transcription factors. The histone acetyltransferase (HAT) catalytic domain of the CBP enzymes shows high level of conservation among different species (Fig 1), as well as among other HATs from A. aegypti (S1A and S1B Fig). AaCBP also contains a bromo domain, which binds acetylated lysines. The conserved domains are connected by long stretches of unstructured linkers.

Fig 1. Overview of AaCBP protein domain conservation.

Fig 1

Schematic representation of the full-length AaCBP protein (XP_011493407.2), depicting the conserved functional domains: the TAZ domain (orange), KIX domain (pink), Bromo domain (blue), HAT domain (purple) and CREB domain (green). The full-length CBP from Homo sapiens (NP_001420.2), Apis melifera (XP_026294861.1) and Drosophila melanogaster (AAB53050.1) are also shown for comparison. The percentages of similarity of the CBP-HAT domains are shown within the purple boxes.

Because transcriptional coactivator complexes are intricate structures composed of multiple subunits, and because cooperative assembly of histone acetyltransferases is a rate-limiting step in transcription activation, we aimed to identify other HATs in A. aegypti. By searching the latest A. aegypti proteome and functional annotations (see Methods), we identified six other putative HATs, all containing the conserved HAT functional domain (S1A and S1B Fig).

ZIKV infection modulates the expression and activity of AaCBP

Mosquitoes naturally acquire viral infections when they feed on blood. In this context, it is well established that in addition to extensive genome-wide transcriptional modulation in mosquitoes after blood meals [31], viral infections can also modulate host cell gene expression and influence cellular function. Thus, we first evaluated whether a blood meal was able to modulate the expression of AaCBP in the midgut or fat body (S2 Fig). We showed that the expression of AaCBP in both tissues was significantly upregulated by blood meal, reaching its peak of transcription at 24 and 48 h after feeding (S2A and S2B Fig). The stimulated transcription of AaCBP was especially dramatic in the fat body 48 h after a blood meal (S2B Fig). Because we were particularly interested in evaluating the functional role of AaCBP in ZIKV infection, we were forced to change our virus infection approach. As an alternative to feeding infected blood, mosquitoes were infected with ZIKV by intrathoracic injections. Importantly, we showed that ZIKV infection was also able to upregulate the expression of AaCBP in the midgut or fat body, reaching its peak transcription at four days after infection (Fig 2A and 2B). However, we did not see upregulation when we assayed head and thorax (Fig 2C). A similar phenomenon was observed when we used Aag2 cells, where the expression of AaCBP reached their peaks at 6 and 15 h, post infection, respectively (Fig 2D). Importantly, we confirmed that the increase in mRNA expression correlated with the increase in AaCBP acetylation activity (Fig 2E and 2F). Of note, ZIKV infection enhanced the acetylation activity of AaCBP toward lysine 27 of histone H3 (H3K27ac), but not toward lysine 9 of histone H3 (H3K9ac), (Fig 2F). Considering that H3K27 is the main target substrate for CBP enzymes [32], and that H3K9 is the main substrate of Gcn5 HAT [33] (also present in A. aegypti; see S1A and S1B Fig), these results point to a specific enhancement of AaCBP by ZIKV infection.

Fig 2. ZIKV infection modulates the expression and activity of AaCBP.

Fig 2

Aag2 cells as well as mosquitoes were infected with ZIKV at a MOI of 2.0, and 60 PFUs, respectively. Mosquito infections were performed by intrathoracic injections. The expression of AaCBP in mosquitoes (A-C) or Aag2 cells (D) was measured by qRT-PCR on the indicated tissues and days post infection. E and F. Western blot of total protein extract from Aag2 cells infected with ZIKV virus, or mock-infected at different time points. Monoclonal antibodies against H3K9ac, H3K27ac, H3 panacetylated (Panac), or H3 (as a loading control) were used. The intensity of the bands was quantified by densitometry analysis plotted as a graph using ImageJ (NIH Software). Western blotting was performed on 3 independent biological replicates and one representative is shown here. Error bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05; **, p <0.01.

AaCBP plays a role in the defense and survival of ZIKV-infected mosquitoes

To investigate the role of AaCBP in the lifespan of mosquitoes infected with ZIKV, we knocked down the AaCBP gene two days before infection and followed their survival rates for 20 days, on a daily basis (Fig 3A). It is important to emphasize that it was mandatory to use intrathoracic injections for ZIKV infection in these experiments. This was due to the fact that AaCBP gene silencing was not successfully achieved when a blood meal was utilized (S3 Fig; of note, we showed that the expression of AaCBP is upregulated by blood meal, and thus, this likely counteracted the effects of the silencing). The levels of silencing of AaCBP at Day 2 post dsRNA injections were 50% in the midgut and 80% in the fat body, as judged by its mRNA expression and activity (S4A and S4B Fig). Mock-infected mosquitoes that received dsLuc injections, showed a high rate of survival until Day 15, when survival started to decline, most likely due to normal aging (Fig 3B, black lines. Mock-infected mosquitoes that were silenced for AaCBP revealed a mortality rate of 60% at Day 20 (Fig 3B, blue line), a similar pattern observed for ZIKV-infected mosquitoes that received injections with the dsLuc control (Fig 3B, green line). Importantly, mosquitoes that were silenced for AaCBP followed by ZIKV infections showed a much higher mortality rate than all other groups (Fig 3B, orange line). These results indicate that even a partial deletion of AaCBP is enough to disrupt the homeostasis of the mosquito (Fig 3B blue line), and when virus infection occurs, the lack of AaCBP becomes enormously detrimental.

Fig 3. AaCBP gene knockdown compromises the lifespan of ZIKV-infected mosquitoes.

Fig 3

A. One hundred adult female mosquitoes were injected intrathoracically with double-stranded RNAs for the AaCBP or Luc control gene, two days before Zika infections (by intrathoracic injections). Survival of the mosquitoes was monitored on a daily basis, until Day 20. B. The experiments were conducted in five 500ml cups, each containing 20 mosquitoes. Dead mosquitoes were removed on a daily basis. The survival curves were repeated at least five times. Survival curve of mosquitoes that were not injected, mock-infected mosquitoes that were injected with for dsLuc or dsAaCBP, and mosquitoes that were injected with for dsLuc or dsAaCBP and infected with ZIKV. C and D. Midguts or fat bodies from silenced mosquitoes were assessed for virus infection intensity at different time points. Each dot or square represents the mean plaque-forming units (PFUs) per two midguts or fat bodies (totaling 20 midguts or fat bodies examined) from five independent experiments. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05; **, p < 0.01.

To correlate the lack of AaCBP and mortality, with an increase in viral loads, we determined viral titers over time in the midguts and fat bodies (Fig 3C and 3D, respectively). We saw that AaCBP-silenced mosquitoes did not efficiently fight virus infections in these tissues that are expected to mount strong immune responses against viruses [4,5,34]. In this respect, our data reconfirmed that antimicrobial peptides (AMPs) are highly upregulated by ZIKV infections in the midgut (S5 Fig).

AaCBP regulates the expression of antiviral immune-response genes

Because AaCBP-silenced mosquitoes revealed higher viral loads than control-silenced mosquitoes (Fig 3C and 3D), and knowing the transcriptional coactivator role of CBP enzymes, we questioned whether AaCBP knockdown could have affected the transcription of antiviral immune-response genes. Indeed, qRT-PCR analysis showed that the lack of AaCBP led to downregulation of the immune-response genes cecropin D, cecropin G, defensin A and defensin C, in the midgut or fat body (Fig 4A and 4B, respectively). In addition, the vago 2 gene was also downregulated in the fat body (Fig 4B), but not defensin A. Interestingly, vago 2 has been shown to be upregulated in A. aegypti larvae exposed to dengue virus (DENV) [35], as well as in A. albopictus cells infected with DENV [36]. Of note, two members of the RNAi (siRNA) pathway, dicer 2 and argonaute 2 were not modulated in either tissue of AaCBP-silenced mosquitoes (Fig 4A and 4B). Importantly, downregulation of antiviral immune-response genes persisted until at least day five-post infection, which matched the kinetics of AaCBP-mediated gene silencing (S6A–S6D Fig).

Fig 4. The expression of antiviral immune-response genes is reduced in ZIKV-infected mosquitoes silenced for AaCBP.

Fig 4

A and B. Fifty adult female mosquitoes were silenced for AaCBP, and after two days, mosquitoes were infected with ZIKV. The expression of cecropin D (CEC D), cecropin G (CEC G), attacin (ATTAC), serpin (SERP), defensin A (DEF A), defensin C (DEF C), dicer 2, argonaute 2 (AGO2) and vago 2 in the midgut (A) or fat body (B) was measured by qRT-PCR four days post infection. Silencing levels of AaCBP in both tissues are shown (first bars in Panels A and B). qRT-PCR was performed from 3 independent biological replicates. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05.

There is always the possibility that the observed dysregulation of immune genes in AaCBP-silenced mosquitoes was accompanied by intestinal dysbiosis. To rule out this possibility, we carried out genomic DNA qPCR to quantify the total bacterial or fungal load in the midguts of AaCBP-silenced or control mosquitoes (S7 Fig). We showed that the midgut microbiota was not affected by AaCBP-silencing (S7 Fig).

Histone hyperacetylation induces mosquito immune responses and suppression of ZIKV infection

Histone deacetylation by histone deacetylases (HDACs) is involved in chromatin compaction and gene repression [37]. Inhibition of HDACs by sodium butyrate (NaB) leads to histone hyperacetylation and potent gene activation [38]. We treated mosquitoes with NaB and showed a significant increase in H3K27 and H3K9 acetylation (S8A and S8B Fig). An increase in total histone H3 acetylation could not be observed (S8A and S8B Fig). Although we do not have a clear interpretation of this result, we can only speculate at this stage that the levels of the specific acetylated lysines can vary among the different tissues of the whole mosquito, since we are analyzing specific marks (K9 or K27) unlike the other five H3 lysine marks (K4, K14, K18, K23 AND K56).

We next investigated the effect of NaB in ZIKV-infected mosquitoes (Fig 5) and showed that the expression levels of defensin A, defensin C and cecropin D were significantly increased (Fig 5A–5C). Importantly, the effect of NaB treatment consistently showed a significant reduction in viral loads (Fig 5D and 5E). It is important to emphasize that NaB treatment leads to hyperacetylation of all histones. In this respect, we observed hyperacetylation of H3K27, (a CBP target) and H3K9 (a Gcn5 target). However, the acetylation levels of H3K27 were higher than those of H3K9 (S8A and S8B Fig), likely due to the specific enhancement of AaCBP gene expression and activity by ZIKV infection (Fig 2), which was not observed for AaGcn5 activity (Fig 2F, H3K9ac panel). All together, these data suggest a direct correlation between high chromatin decompaction, AMPs overexpression, and strong immunity.

Fig 5. Histone hyperacetylation in ZIKV-infected mosquitoes induces immune responses and suppresses virus infection.

Fig 5

A-E. Fifty adult female mosquitoes were infected with ZIKV and treated with 0.5 M NaB for three days. A-C. The expression of defensin A, defensin C, and cecropin D, in whole mosquito tissues was measured by qRT-PCR. qRT-PCR was performed on 3 independent biological replicates. D and E. Midguts or fat bodies from ZIKV-infected mosquitoes treated with NaB were assessed for infection intensity at 7 or 15 days post infection. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05; **, p < 0.01.

Discussion

Upon pathogen detection, the innate immune system must be able to mount a robust and quick response, but equally important is the need to rein in the cytotoxic effects of such a response. Immune-response genes are maintained in a silent, yet poised, state that can be readily induced in response to a particular pathogen, and this characteristic pattern is achieved through the action of two elements: the activation of transcription factors and the modulation of the chromatin environment at gene promoters. Although the activation route of the immune pathways against viruses is relatively well known in A. aegypti, our work is the first to attempt to explore the role of chromatin structure in this process.

The regulation of cellular functions by gene activation is accomplished partially by acetylation of histone proteins to open the chromatin conformation, and strikingly, CBP histone acetyltransferase activity always plays a role in this process [39]. One example of signals that ultimately use CBP enzymes as transcriptional coregulators includes the NF-kB signaling [40].

CBP genes are conserved in a variety of multicellular organisms, from worms to humans and play a central role in coordinating and integrating multiple cell signal-dependent events. In this regard, the A. aegypti CBP shows a high degree of homology, notably within the functional domains, with well-characterized human and fly enzymes (Fig 1). Therefore, one might anticipate that AaCBP functions as a transcriptional coactivator in a variety of physiological processes of the mosquito, including innate immunity.

The Toll pathway is an NF-kB pathway that is suggested to play a role in immunity in mosquitoes [4,5,34]. Gene expression analysis of ZIKV-infected mosquitoes has shown that Toll pathway-related genes are upregulated in ZIKV infection when compared to other immune pathways [4]. Here, we showed that the expression of AaCBP is also upregulated upon ZIKV infection and that CBP-dependent histone acetylation enables the mosquito to fight viral infections. Although it is not yet clear how AaCBP could limit virus replication, one could envision a molecular role where a particular transcription factor (for example, Rel1) would recruit AaCBP to immune-related gene (for example, AMP genes) promoters and/or enhancers and acetylate histones (for example, H3K27), culminating with chromatin decompaction and gene activation (depicted in our hypothetical model in Fig 6). Indeed, we experimentally demonstrated in part that our model might be correct: 1. We showed that ZIKV-infection potentiates AaCBP-mediated H3K27 acetylation (a mark of gene activation); 2. The lack of AaCBP leads to downregulation of immune-related genes, higher loads of ZIKV virus, and lower rates of mosquito survival; 3. An inverse phenotype was obtained under H3K27 hyperacetylation. Nevertheless, the AaCBP-bound transcription factor in this signaling pathway has yet to be determined.

Fig 6. Proposed model for the role of AaCBP-mediated histone acetylation in suppressing ZIKV infection in A. aegypti.

Fig 6

Zika infection signals the functional activation of transcription factors (TF) that translocate to the nucleus where they activate the expression of immune-related genes to fight the virus infection. Chromatin decondensation is mandatory for gene activation. Therefore, AaCBP is likely recruited by TFs to target promoters, where it exerts its histone acetyltransferase activity, leading to an open state of the chromatin at these promoter sites.

Host-pathogen interactions provide a highly plastic and dynamic biological system. To cope with the selective constraints imposed by their hosts, many pathogens have evolved unparalleled levels of phenotypic plasticity in their life history traits [41]. Likewise, the host phenotype is drastically and rapidly altered by the presence of a pathogen [42,43]. One important example of these alterations is the manipulative strategy of the pathogen aimed at maximizing its survival and transmission, and one obvious target is the host’s immune system. In recent years, the epigenetic modulation of the host’s transcriptional program linked to host defense has emerged as a relatively common occurrence of pathogenic viral infections [44]. Interestingly and importantly, our data reveal that chromatin remodeling by histone acetylation contributes to establishing a resistance in ZIKV-infected A. aegypti. In this respect, we showed that ZIKV-infected wild type A. aegypti resisted to infections better than ZIKV-infected-AaCBP-silenced mosquitoes, whose survival was drastically compromised (Fig 3B). Thus, our data point to an important role of AaCBP in maintaining A. aegypti homeostasis through fine-tuning the transcriptional control of immune genes.

Overall, we believe that the disruption of AaCBP would have an effect on immune-gene regulation by compromising its physical interaction with transcription factors and the transcriptional machinery, and/or its enzymatic activity, leading to a significant increase in viral loads. Nevertheless, the precise mechanism that led to the enhancement of ZIKV replication after AaCBP knockdown is still missing. This is particularly true when we compare the kinetics between ZIKV loads (Fig 3C and 3D) and mRNA transcription of AMPs (S6A–S6D Fig.).

Our work has attempted to explore the epigenetic nature of virus-vector interactions. We have focused on epigenetic events that initiate changes in the vector nucleus, likely involving the cooperation between transcription factors and chromatin modifiers to integrate and initiate genomic events, which culminate with the limitation of Zika virus replication.

Supporting information

S1 Table. List of primers used in this study.

(XLSX)

S1 Fig. Protein domain and alignment of putative histone acetyltransferases from A. aegypti.

A. Genome identification of putative lysine acetyltransferase (KAT) homologs from A. aegypti, AaHAT1 (XP_001651817.1), AaTip60 (XP_021706851.1), AaMOF1 (XP_001658578.2), AaMOF2 (XP_021711683.1), HBO1 (XP_021701314.1), GCN5 (XP_0016566424.1) and AaCBP itself. Functional domains are indicated above each box. B. Protein sequence alignment of the HAT domains from the putative A. aegypti KATs. Amino acids in red show identity or conservation among all 7 HAT domains.

(TIF)

S2 Fig. Blood meal upregulates the expression of AaCBP.

Mosquitoes were fed with blood over different time courses and mRNA quantification by qPCR was performed in the midgut or fat body. The results in A and B are pools of at least 3 independent experiments, plotted using samples from sugar-fed mosquitoes as reference. Error bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05; **, p < 0.01; ***, p <0.001.

(TIF)

S3 Fig. Silencing levels of AaCBP after a sugar or blood meal.

Two days after feeding, the AaCBP gene was knocked down and its expression was measured two days after silencing. qRT-PCR was performed from 3 independent biological replicates. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p < 0.05.

(TIF)

S4 Fig. Silencing of AaCBP in adult female mosquitoes by intrathoracic injections of dsRNAs.

The expression (A) and activity (B) of AaCBP were used to evaluate the efficiency of gene knockdown. A. The mRNA levels of AaCBP in the midgut, ovary, fat body or whole mosquito were quantified by qRT-PCR at 48 h postinjection with dsCBP or dsLuc. Silencing level was determined by the ratio between mRNA levels of AaCBP-silenced versus dsLuc-injected mosquitoes. B. Western blot of 10 μg of total protein extract of dsCBP- or dsLuc-injected-mosquitoes. Monoclonal antibodies against acetylated- or nonacetylated histone H3 (loading control) are indicated. The intensity of the bands was quantified by densitometry analysis plotted as a graph using ImageJ (NIH Software). Western blotting was performed on 3 independent biological replicates and one representative is shown here. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05.

(TIF)

S5 Fig. ZIKV infection upregulates the expression of antimicrobial peptides in the midgut of A. aegypti.

Fifty adult female mosquitoes were infected with ZIKV after feeding on infected blood and the expression of cecropin D (CEC D), cecropin G (CEC G), attacin (ATTC), defensin C (DEF C) and serpin (SERP) was measured by qRT-PCR four days post infection. qRT-PCR was performed from 3 independent biological replicates. Bars indicate the standard error of the mean.

(TIF)

S6 Fig. Silencing of AaCBP is sustainable and efficient until five days of ZIKV infection.

Fifty adult female mosquitoes were infected with ZIKV for 1, 5, 7 or 10 days and the expression levels of AaCBP, cecropin G, defensin C or vago 2 in the fat body were measured by qRT-PCR on 3 independent biological replicates. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05.

(TIF)

S7 Fig. Intestinal microbiota load by genomic DNA qPCR.

Fifty nanograms of genomic DNA from a pool of 15 midguts from silenced female mosquitoes was used as a template. Amplifications were carried out using specific primers for the ribosomal genes 16S and 18S, for bacteria or fungi, respectively. The A. aegypti ribosomal protein 49 gene (Rp49) was used as an endogenous control. Quantifications were carried out using the comparative Ct method. Genomic DNA qPCR was performed from 6 independent biological replicates. Bars indicate the standard error of the mean.

(TIF)

S8 Fig. Histone deacetylase inhibition leads to histone hyperacetylation in A. aegypti.

Ten adult female mosquitoes were intrathoracically injected with PBS, 0.25 M (17.5 pmol), or 0.5 M (35 pmol) of sodium butyrate (NaB). Four days after treatment, 10 μg of total protein extract from 10 mosquitoes was used for histone acetylation analysis. Western blotting with monoclonal antibodies against H3K9ac, H3K27ac, H3 panacetylated, or H3 (as loading control) was performed. The intensity of the bands was quantified by densitometry (lower panel) analysis plotted as a graph using ImageJ (NIH Software). Western blotting was performed on 3 independent biological replicates and one representative is shown in panel A. Error bars in Panel B indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05.

(TIF)

Acknowledgments

We thank Jaciara Miranda Freire for running the insectary and excellent technical assistance with mosquito rearing and Dr. Ana Cristina Bahia Nascimento (IBCCF–UFRJ) for productive discussions.

Data Availability

All relevant data are within the manuscript and its Supporting Information files.

Funding Statement

This work was supported by grants from Conselho Nacional de Desenvolvimento Científico e Tecnológico (470099/20143) and Fundação Carlos Chagas Filho de Amparo a Pesquisa do Estado do Rio de Janeiro (E-26/202990/2015) to MRF. PLO was supported by Instituto Nacional de Ciência e Tecnologia em Entomologia Molecular (573959/2008-0). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.

References

  • 1.Weaver SC, Costa F, Garcia-Blanco MA, Ko AI, Ribeiro GS, Saade G, et al. Zika virus: History, emergence, biology, and prospects for control. Antiviral Res. 2016;130: 69–80. doi: 10.1016/j.antiviral.2016.03.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Gómez-Díaz E, Jordà M, Peinado MA, Rivero A. Epigenetics of Host-Pathogen Interactions: The Road Ahead and the Road Behind. PLoS Pathog. 2012;8. doi: 10.1371/journal.ppat.1003007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Ruiz JL, Yerbanga RS, Lefèvre T, Ouedraogo JB, Corces VG, Gómez-Díaz E. Chromatin changes in Anopheles gambiae induced by Plasmodium falciparum infection. Epigenetics Chromatin. 2019;12: 5. doi: 10.1186/s13072-018-0250-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Angleró-Rodríguez YI, MacLeod HJ, Kang S, Carlson JS, Jupatanakul N, Dimopoulos G. Aedes aegypti Molecular Responses to Zika Virus: Modulation of Infection by the Toll and Jak/Stat Immune Pathways and Virus Host Factors. Front Microbiol. 2017;8: 2050. doi: 10.3389/fmicb.2017.02050 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Tikhe C V, Dimopoulos G. Mosquito antiviral immune pathways. Dev Comp Immunol. 2021;116: 103964. doi: 10.1016/j.dci.2020.103964 [DOI] [PubMed] [Google Scholar]
  • 6.Asad S, Parry R, Asgari S. Upregulation of Aedes aegypti Vago1 by Wolbachia and its effect on dengue virus replication. Insect Biochem Mol Biol. 2018;92: 45–52. doi: 10.1016/j.ibmb.2017.11.008 [DOI] [PubMed] [Google Scholar]
  • 7.Barrera LO, Ren B. The transcriptional regulatory code of eukaryotic cells–insights from genome-wide analysis of chromatin organization and transcription factor binding. Curr Opin Cell Biol. 2006;18: 291–298. doi: 10.1016/j.ceb.2006.04.002 [DOI] [PubMed] [Google Scholar]
  • 8.Chan HM, Thangue NB La. p300 / CBP proteins: HATs for transcriptional bridges and scaffolds. 2001; 114(Pt 13):2363–2373. doi: 10.1242/jcs.114.13.2363 [DOI] [PubMed] [Google Scholar]
  • 9.Revilla Y, Granja AG. Viral mechanisms involved in the transcriptional CBP/p300 regulation of inflammatory and immune responses. Crit Rev Immunol. 2009;29: 131–154. doi: 10.1615/critrevimmunol.v29.i2.30 [DOI] [PubMed] [Google Scholar]
  • 10.Fang F, Xu Y, Chew K-K, Chen X, Ng H-H, Matsudaira P. Coactivators p300 and CBP maintain the identity of mouse embryonic stem cells by mediating long-range chromatin structure. Stem Cells. 2014;32: 1805–1816. doi: 10.1002/stem.1705 [DOI] [PubMed] [Google Scholar]
  • 11.Bantignies F, Goodman RH, Smolik SM. The interaction between the coactivator dCBP and Modulo, a chromatin-associated factor, affects segmentation and melanotic tumor formation in Drosophila. Proc Natl Acad Sci U S A. 2002;99: 2895–2900. doi: 10.1073/pnas.052509799 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Fauquier L, Azzag K, Parra MAM, Quillien A, Boulet M, Diouf S, et al. CBP and P300 regulate distinct gene networks required for human primary myoblast differentiation and muscle integrity. Sci Rep. 2018;8: 12629. doi: 10.1038/s41598-018-31102-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Iyer NG, Ozdag H, Caldas C. p300/CBP and cancer. Oncogene. 2004;23: 4225–4231. doi: 10.1038/sj.onc.1207118 [DOI] [PubMed] [Google Scholar]
  • 14.Kirfel P, Vilcinskas A, Skaljac M. Lysine Acetyltransferase p300/CBP Plays an Important Role in Reproduction, Embryogenesis and Longevity of the Pea Aphid Acyrthosiphon pisum. Insects. 2020;11. doi: 10.3390/insects11050265 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Li K-L, Zhang L, Yang X-M, Fang Q, Yin X-F, Wei H-M, et al. Histone acetyltransferase CBP-related H3K23 acetylation contributes to courtship learning in Drosophila. BMC Dev Biol. 2018;18: 20. doi: 10.1186/s12861-018-0179-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Sedkov Y, Cho E, Petruk S, Cherbas L, Smith ST, Jones RS, et al. Methylation at lysine 4 of histone H3 in ecdysone-dependent development of Drosophila. Nature. 2003;426: 78–83. doi: 10.1038/nature02080 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Tie F, Banerjee R, Stratton CA, Prasad-Sinha J, Stepanik V, Zlobin A, et al. CBP-mediated acetylation of histone H3 lysine 27 antagonizes Drosophila Polycomb silencing. Development. 2009;136: 3131–3141. doi: 10.1242/dev.037127 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Donald CL, Brennan B, Cumberworth SL, Rezelj V V., Clark JJ, Cordeiro MT, et al. Full Genome Sequence and sfRNA Interferon Antagonist Activity of Zika Virus from Recife, Brazil. PLoS Negl Trop Dis. 2016;10: 1–20. doi: 10.1371/journal.pntd.0005048 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Göertz GP, van Bree JWM, Hiralal A, Fernhout BM, Steffens C, Boeren S, et al. Subgenomic flavivirus RNA binds the mosquito DEAD/H-box helicase ME31B and determines Zika virus transmission by Aedes aegypti. Proc Natl Acad Sci U S A. 2019;116: 19136–19144. doi: 10.1073/pnas.1905617116 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Sim S, Jupatanakul N, Ramirez JL, Kang S, Romero-Vivas CM, Mohammed H, et al. Transcriptomic profiling of diverse Aedes aegypti strains reveals increased basal-level immune activation in dengue virus-refractory populations and identifies novel virus-vector molecular interactions. PLoS Negl Trop Dis. 2013;7: e2295. doi: 10.1371/journal.pntd.0002295 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Ramirez JL, Souza-Neto J, Torres Cosme R, Rovira J, Ortiz A, Pascale JM, et al. Reciprocal Tripartite Interactions between the Aedes aegypti Midgut Microbiota, Innate Immune System and Dengue Virus Influences Vector Competence. PLoS Negl Trop Dis. 2012;6: 1–11. doi: 10.1371/journal.pntd.0001561 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods. 2001;25: 402–408. doi: 10.1006/meth.2001.1262 [DOI] [PubMed] [Google Scholar]
  • 23.Gentile C, Lima JBP, Peixoto AA. Isolation of a fragment homologous to the rp49 constitutive gene of Drosophila in the Neotropical malaria vector Anopheles aquasalis (Diptera: Culicidae). Mem Inst Oswaldo Cruz. 2005;100: 545–547. doi: 10.1590/s0074-02762005000600008 [DOI] [PubMed] [Google Scholar]
  • 24.Coutinho Carneiro V, de Abreu da Silva IC, Amaral MS, Pereira ASA, Silveira GO, Pires D da S, et al. Pharmacological inhibition of lysine-specific demethylase 1 (LSD1) induces global transcriptional deregulation and ultrastructural alterations that impair viability in Schistosoma mansoni. PLoS Negl Trop Dis. 2020;14: e0008332. doi: 10.1371/journal.pntd.0008332 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Ribeiro FS, de Abreu da Silva IC, Carneiro VC, Belgrano F dos S, Mohana-Borges R, de Andrade Rosa I, et al. The dengue vector Aedes aegypti contains a functional high mobility group box 1 (HMGB1) protein with a unique regulatory C-terminus. PLoS One. 2012;7: e40192. doi: 10.1371/journal.pone.0040192 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Altschul SF, Madden TL, Schäffer AA, Zhang J, Zhang Z, Miller W, et al. Gapped BLAST and PSI-BLAST: a new generation of protein database search programs. Nucleic Acids Res. 1997;25: 3389–3402. doi: 10.1093/nar/25.17.3389 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Feller C, Forné I, Imhof A, Becker PB. Global and specific responses of the histone acetylome to systematic perturbation. Mol Cell. 2015;57: 559–571. doi: 10.1016/j.molcel.2014.12.008 [DOI] [PubMed] [Google Scholar]
  • 28.Ren J, Wen L, Gao X, Jin C, Xue Y, Yao X. DOG 1.0: illustrator of protein domain structures. Cell research. England; 2009. pp. 271–273. doi: 10.1038/cr.2009.6 [DOI] [PubMed] [Google Scholar]
  • 29.Thompson JD, Higgins DG, Gibson TJ. CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Res. 1994;22: 4673–4680. doi: 10.1093/nar/22.22.4673 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Liu X, Wang L, Zhao K, Thompson PR, Hwang Y, Marmorstein R, et al. The structural basis of protein acetylation by the p300/CBP transcriptional coactivator. Nature. 2008;451: 846–850. doi: 10.1038/nature06546 [DOI] [PubMed] [Google Scholar]
  • 31.Bonizzoni M, Dunn WA, Campbell CL, Olson KE, Dimon MT, Marinotti O, et al. RNA-seq analyses of blood-induced changes in gene expression in the mosquito vector species, Aedes aegypti. BMC Genomics. 2011;12: 82. doi: 10.1186/1471-2164-12-82 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Raisner R, Kharbanda S, Jin L, Jeng E, Chan E, Merchant M, et al. Enhancer Activity Requires CBP/P300 Bromodomain-Dependent Histone H3K27 Acetylation. Cell Rep. 2018;24: 1722–1729. doi: 10.1016/j.celrep.2018.07.041 [DOI] [PubMed] [Google Scholar]
  • 33.Karmodiya K, Krebs AR, Oulad-Abdelghani M, Kimura H, Tora L. H3K9 and H3K14 acetylation co-occur at many gene regulatory elements, while H3K14ac marks a subset of inactive inducible promoters in mouse embryonic stem cells. BMC Genomics. 2012;13: 424. doi: 10.1186/1471-2164-13-424 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Cheng G, Liu Y, Wang P, Xiao X. Mosquito Defense Strategies against Viral Infection. Trends Parasitol. 2016. doi: 10.1016/j.pt.2015.09.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Vargas V, Cime-Castillo J, Lanz-Mendoza H. Immune priming with inactive dengue virus during the larval stage of Aedes aegypti protects against the infection in adult mosquitoes. Sci Rep. 2020;10: 6723. doi: 10.1038/s41598-020-63402-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Paradkar PN, Duchemin J, Voysey R, Walker PJ. Dicer-2-Dependent Activation of Culex Vago Occurs via the TRAF-Rel2 Signaling Pathway. 2014;8. doi: 10.1371/journal.pntd.0002823 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Shahbazian MD, Grunstein M. Functions of site-specific histone acetylation and deacetylation. Annu Rev Biochem. 2007;76: 75–100. doi: 10.1146/annurev.biochem.76.052705.162114 [DOI] [PubMed] [Google Scholar]
  • 38.Kurdistani SK, Grunstein M. Histone acetylation and deacetylation in yeast. Nat Rev Mol Cell Biol. 2003;4: 276–284. doi: 10.1038/nrm1075 [DOI] [PubMed] [Google Scholar]
  • 39.Dancy BM, Cole PA. Protein lysine acetylation by p300/CBP. Chem Rev. 2015;115: 2419–2452. doi: 10.1021/cr500452k [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Mukherjee SP, Behar M, Birnbaum HA, Hoffmann A, Wright PE, Ghosh G. Analysis of the RelA:CBP/p300 interaction reveals its involvement in NF-κB-driven transcription. PLoS Biol. 2013;11: e1001647. doi: 10.1371/journal.pbio.1001647 [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Reece SE, Ramiro RS, Nussey DH. Plastic parasites: sophisticated strategies for survival and reproduction? Evol Apl. 2009; 11–23. doi: 10.1111/j.1752-4571.2008.00060.x [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Edelaar P, Bonduriansky R, Charmantier A, Danchin E, Pujol B. Response to Kalchhauser et al.: Inherited Gene Regulation Is not Enough to Understand Nongenetic Inheritance. Trends in ecology & evolution. England; 2021. pp. 475–476. doi: 10.1016/j.tree.2021.03.004 [DOI] [PubMed] [Google Scholar]
  • 43.Rando OJ, Verstrepen KJ. Timescales of genetic and epigenetic inheritance. Cell. 2007;128: 655–668. doi: 10.1016/j.cell.2007.01.023 [DOI] [PubMed] [Google Scholar]
  • 44.Paschos K, Allday MJ. Epigenetic reprogramming of host genes in viral and microbial pathogenesis. Trends Microbiol. 2010;18: 439–447. doi: 10.1016/j.tim.2010.07.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
PLoS Negl Trop Dis. doi: 10.1371/journal.pntd.0010559.r001

Decision Letter 0

Philippe Desprès, Jason L Rasgon

21 Jan 2022

Dear Dr Fantappié,

Thank you very much for submitting your manuscript "Zika virus infection drives epigenetic modulation of immunity by the histone acetyltransferase CBP of Aedes aegypti" for consideration at PLOS Neglected Tropical Diseases. As with all papers reviewed by the journal, your manuscript was reviewed by members of the editorial board and by several independent reviewers. In light of the reviews (below this email), we would like to invite the resubmission of a significantly-revised version that takes into account the reviewers' comments.

We cannot make any decision about publication until we have seen the revised manuscript and your response to the reviewers' comments. Your revised manuscript is also likely to be sent to reviewers for further evaluation.

When you are ready to resubmit, please upload the following:

[1] A letter containing a detailed list of your responses to the review comments and a description of the changes you have made in the manuscript. Please note while forming your response, if your article is accepted, you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out.

[2] Two versions of the revised manuscript: one with either highlights or tracked changes denoting where the text has been changed; the other a clean version (uploaded as the manuscript file).

Important additional instructions are given below your reviewer comments.

Please prepare and submit your revised manuscript within 60 days. If you anticipate any delay, please let us know the expected resubmission date by replying to this email. Please note that revised manuscripts received after the 60-day due date may require evaluation and peer review similar to newly submitted manuscripts.

Thank you again for your submission. We hope that our editorial process has been constructive so far, and we welcome your feedback at any time. Please don't hesitate to contact us if you have any questions or comments.

Sincerely,

Jason L. Rasgon

Associate Editor

PLOS Neglected Tropical Diseases

Philippe Desprès

Deputy Editor

PLOS Neglected Tropical Diseases

***********************

Reviewer's Responses to Questions

Key Review Criteria Required for Acceptance?

As you describe the new analyses required for acceptance, please consider the following:

Methods

-Are the objectives of the study clearly articulated with a clear testable hypothesis stated?

-Is the study design appropriate to address the stated objectives?

-Is the population clearly described and appropriate for the hypothesis being tested?

-Is the sample size sufficient to ensure adequate power to address the hypothesis being tested?

-Were correct statistical analysis used to support conclusions?

-Are there concerns about ethical or regulatory requirements being met?

Reviewer #1: This study fulfills above mentioned requirements. Please see the summary and general comments section for a detailed review comments.

Reviewer #2: Critical details regarding some methods are missing, including sample sizes, number of replications and statistical tests.

--------------------

Results

-Does the analysis presented match the analysis plan?

-Are the results clearly and completely presented?

-Are the figures (Tables, Images) of sufficient quality for clarity?

Reviewer #1: Yes.

Please see the summary and general comments section for a detailed review comments.

Reviewer #2: Results are clearly presented and figures are good quality

--------------------

Conclusions

-Are the conclusions supported by the data presented?

-Are the limitations of analysis clearly described?

-Do the authors discuss how these data can be helpful to advance our understanding of the topic under study?

-Is public health relevance addressed?

Reviewer #1: Yes.

Please see the summary and general comments section for a detailed review comments.

Reviewer #2: The conclusions are only partially supported by the data. Some contradictions are not discussed, nor are alternative explanations developed and discussed.

--------------------

Editorial and Data Presentation Modifications?

Use this section for editorial suggestions as well as relatively minor modifications of existing data that would enhance clarity. If the only modifications needed are minor and/or editorial, you may wish to recommend “Minor Revision” or “Accept”.

Reviewer #1: Accept with minor revision

Reviewer #2: (No Response)

--------------------

Summary and General Comments

Use this section to provide overall comments, discuss strengths/weaknesses of the study, novelty, significance, general execution and scholarship. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. If requesting major revision, please articulate the new experiments that are needed.

Reviewer #1: This exciting manuscript investigates the epigenetic modulation of the Zika virus in Ae. aegypti mosquitoes. Authors have performed bioinformatics characterization of CREB-binding protein (CBP) orthologue in Aedes aegypti (AaCBP), and expression pattern of AaCBP in different tissues of ZIKV infected Aedes aegypti mosquitoes. The CBP promotes transcription through chromatin remodeling, acetylation of proteins, and recruitment of the basal transcription machinery. Using pharmacological activator and genetic knockdown strategies, authors have demonstrated a viable model explaining how Zika virus-mediated acetylation of histone proteins moderated chromatin conformation that modulates innate immune response of Aedes aegypti and maintains homeostasis and pathological balance to tolerate persistent infection and develop efficient antiviral strategies to restrict viral replication to non-pathogenic levels.

Although the experiments are well planned and executed, the following minor but important clarifications/additions would help improve the manuscript:

1. In Figure 4 the authors identified three immune genes that are down-regulated in response to dsRNA KD of AaCBP, namely Cecropin D, Cecropin G, and Defensin C. However, in Figure 5 they did not include Cecropin G for analysis and included Defensin A. An explanation or addition of data for Defensin A in figure 4 would bring more clarity for the readers and consistency to the story.

2. In figure 4 legend, defensin is misspelled. Also, to maintain consistency with the other gene names, please change the first alphabet of gene names ‘dicer’ and ‘argonaute’ into lower case letters.

3. Figure 6 can be presented in a better way e.g., the cytoplasmic and nuclear compartments need to be separated. The mechanistic details as explained in the results section, figure legend and discussion are missing in the figure.

Reviewer #2: Amarante and colleagues present an analysis of the Ae. aegypti transcription factor and histone acyltransferase, AaCBP, in modulating ZIKV infection. AaCBP was induced by ZIKV infection, and knockdown of AaCBP1 decreased the expression of various immune effectors. Additional data after knockdown showed some increases in virus titer in the mosquito and suggest decreases in overall survival. The authors go on to propose a model whereby ZIKV induces AaCBP, which in turn facilitates the activation of immune genes through modifications in chromatin structure. Overall, the work is highly interesting, but there are some gaps and inconsistencies between what the authors report in their figures and their interpretation.

Concerns

For survival curves the authors report “much higher mortality” for dsAsCBP-ZIKV group, but no statistical analysis is presented. The legend states that 100 mosquitoes were used per group, but no indication is given as to whether this experiment was repeated (or how many times), the size/type of the cage these mosquitoes were held in, or whether dead individuals were removed on a daily basis. It is not uncommon for mold to grow on the bodies of dead individuals-in a crowded arena this can exacerbate mortality amongst others in the cage which essentially serves as a feedback loop. This can be mitigated by keeping the mosquitoes in many small groups (5 cups of 20 or even better 10 cups of 10, so an artifactual death spiral would only affect a single replicate cup) or by removing dead individuals each day.

The authors show that silencing of AaCBP is over by 10 DPI, and that levels of AMPs and Vago2 are back to normal by 10 DPI (Fig S6). However, differences in ZIKV titer do not emerge until 10 DPI (Fig 3), with no difference at 5 DPI-even though that is when silencing of AaCBP and the AMPs are maximal. This is not consistent with AaCBP or the indicated AMPs acting directly in anti-ZIKV immunity, and this inconsistency is not addressed in the manuscript even though it goes to the heart of the main conclusions.

Likewise in Figure 5, NaB experiments show that AMPs are induced in ZIKV-infected individuals at 3 DPI with reduced virus titer at 7 DPI, but not 15 DPI. Again, this is not consistent with the data presented in Fig3, where the authors report an increase in viral titer at 15 DPI after knockdown of CBP.

In reference to Figure 3, the authors state “We clearly saw that AaCBP-silenced mosquitoes did not efficiently fight virus infections in these tissues that are expected to mount strong immune responses against viruses” – This phrase overstates the findings presented. The authors observed a less than 2-fold increase in virus titer at 10+15 DPI. No difference at all was observed at 5 DPI. This does not mean that these mosquitoes did not mount an immune response, and it is not clear what “efficient” means in this context since that is typically a measure of how much you get out based on how much you put in. Also, the term “clearly” is subjective and should be removed. Given that virus titers in the control group here were about 4 times lower at 15 DPI as compared to the control group in Figure 5 (~2000 PFUs per midgut vs 8000 PFUs per midgut), the authors need to be much more cautious in their interpretations since there is clearly substantial variation between experiments. This variation is never explained at since the authors are injecting virus should not be present (that is the whole point of injections, to make dose uniform).

The authors state that “An increase in total histone H3 acetylation could not be observed (Supplementary Fig 7A and B), which is somehow expected if one considers that a combination of acetylated and nonacetylated H3 might coexist in a specific cell type and/or during a specific period.” – This does not make sense. Acetylation can occur at H3 positions K9, K14, K18, K23 and K27. The authors document what appears to be an increase in acetylation at positions K9 and K27. Total acetylation did not appear to change. Provided the authors data are reliable, the most obvious possibility to check is that acetylation at some of the other positions has decreased. If cell-type specific events were obfuscating total acetylation, they should obscure K9 or K27-specific acetylation as well. Of course, as the authors state, NaB should increase acetylation across all sites, so this is still a bit inconsistent with the authors conclusions.

“The Toll pathway has been shown to play the most important role in controlling ZIKV infections (Angleró-Rodríguez et al., 2017).” -This is not an accurate representation of the findings of Anglero-Rodriguez. In that paper, the authors found the strongest anti-viral response when the Toll pathway was artificially stimulated as compared only with IMD and Jak/Stat. These experiments did not consider any other immune pathway, of which there are still many.

The authors do not adequately consider other hypotheses that could better explain their findings, making it difficult to come to the same conclusions as are reached in the model presented in Figure 6. Two points that need to be further developed (either experimentally or in the discussion or both) are potential affects on the microbiome and the more general role of CBP outside of its role as a HAT. For the former, dysregulation of AMPs may lead to dysbiosis or greater susceptibility to fungal pathogens-maybe this can be compensated for alone but not with the added stress of ZIKV infection. In the latter, dsRNA experiments cannot disentangle the role of CBP in histone acetylation from its more general role as CREB binding protein in RNA transcription.

--------------------

PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: Yes: Sujit Pujhari

Reviewer #2: No

Figure Files:

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email us at figures@plos.org.

Data Requirements:

Please note that, as a condition of publication, PLOS' data policy requires that you make available all data used to draw the conclusions outlined in your manuscript. Data must be deposited in an appropriate repository, included within the body of the manuscript, or uploaded as supporting information. This includes all numerical values that were used to generate graphs, histograms etc.. For an example see here: http://www.plosbiology.org/article/info%3Adoi%2F10.1371%2Fjournal.pbio.1001908#s5.

Reproducibility:

To enhance the reproducibility of your results, we recommend that you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option to publish peer-reviewed clinical study protocols. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

PLoS Negl Trop Dis. doi: 10.1371/journal.pntd.0010559.r003

Decision Letter 1

Philippe Desprès, Jason L Rasgon

31 May 2022

Dear Dr FANTAPPIE,

Thank you very much for submitting your manuscript "Zika virus infection drives epigenetic modulation of immunity by the histone acetyltransferase CBP of Aedes aegypti" for consideration at PLOS Neglected Tropical Diseases. As with all papers reviewed by the journal, your manuscript was reviewed by members of the editorial board and by several independent reviewers. The reviewers appreciated the attention to an important topic. Based on the reviews, we are likely to accept this manuscript for publication, providing that you modify the manuscript according to the review recommendations.

Thank you for your revision. Reviewer 2 still has some issues that need to be addressed (and I believe they can be done quickly). After you address them, I do not anticipate needing to send this out for further review and will process the resubmission with all due speed.

Please prepare and submit your revised manuscript within 30 days. If you anticipate any delay, please let us know the expected resubmission date by replying to this email.

When you are ready to resubmit, please upload the following:

[1] A letter containing a detailed list of your responses to all review comments, and a description of the changes you have made in the manuscript.

Please note while forming your response, if your article is accepted, you may have the opportunity to make the peer review history publicly available. The record will include editor decision letters (with reviews) and your responses to reviewer comments. If eligible, we will contact you to opt in or out

[2] Two versions of the revised manuscript: one with either highlights or tracked changes denoting where the text has been changed; the other a clean version (uploaded as the manuscript file).

Important additional instructions are given below your reviewer comments.

Thank you again for your submission to our journal. We hope that our editorial process has been constructive so far, and we welcome your feedback at any time. Please don't hesitate to contact us if you have any questions or comments.

Sincerely,

Jason L. Rasgon

Associate Editor

PLOS Neglected Tropical Diseases

Philippe Desprès

Deputy Editor

PLOS Neglected Tropical Diseases

***********************

Thank you for your revision. Reviewer 2 still has some issues that need to be addressed (I believe they can be done quickly). After you address them, I do not anticipate needing to send this out for further review.

Reviewer's Responses to Questions

Key Review Criteria Required for Acceptance?

As you describe the new analyses required for acceptance, please consider the following:

Methods

-Are the objectives of the study clearly articulated with a clear testable hypothesis stated?

-Is the study design appropriate to address the stated objectives?

-Is the population clearly described and appropriate for the hypothesis being tested?

-Is the sample size sufficient to ensure adequate power to address the hypothesis being tested?

-Were correct statistical analysis used to support conclusions?

-Are there concerns about ethical or regulatory requirements being met?

Reviewer #1: Yes

Reviewer #2: Some key information on methods are still missing.

--------------------

Results

-Does the analysis presented match the analysis plan?

-Are the results clearly and completely presented?

-Are the figures (Tables, Images) of sufficient quality for clarity?

Reviewer #1: Yes

Reviewer #2: (No Response)

--------------------

Conclusions

-Are the conclusions supported by the data presented?

-Are the limitations of analysis clearly described?

-Do the authors discuss how these data can be helpful to advance our understanding of the topic under study?

-Is public health relevance addressed?

Reviewer #1: Yes

Reviewer #2: (No Response)

--------------------

Editorial and Data Presentation Modifications?

Use this section for editorial suggestions as well as relatively minor modifications of existing data that would enhance clarity. If the only modifications needed are minor and/or editorial, you may wish to recommend “Minor Revision” or “Accept”.

Reviewer #1: (No Response)

Reviewer #2: (No Response)

--------------------

Summary and General Comments

Use this section to provide overall comments, discuss strengths/weaknesses of the study, novelty, significance, general execution and scholarship. You may also include additional comments for the author, including concerns about dual publication, research ethics, or publication ethics. If requesting major revision, please articulate the new experiments that are needed.

Reviewer #1: In this revised manuscript Amarante et al. have satisfactorily addressed the concerns pointed out by the reviewers. Overall, this work is exciting and analyses the role of epigenetic modulations of histone acetyltransferase CBP in Zika virus interaction in Ae. ageypti. The authors propose a model where ZIKV infection activates TF and induces AaCBP resulting in chromatin decondensation and activation of immune genes.

Reviewer #2: Reviewer #2:

The legend states that 100 mosquitoes were used per group, but no indication is given as to

whether this experiment was repeated (or how many times), the size/type of the cage these

mosquitoes were held in,

Authors:

All experiments in this study were repeated at least three times. The survival curves

were repeated 5 times. This information is provided in the legend of Figure 3B (page

28).

New Comment: No information on replicates is present in the legend of Fig3B.

or whether dead individuals were removed on a daily basis. It is not uncommon for mold to

grow on the bodies of dead individuals-in a crowded arena this can exacerbate mortality

amongst others in the cage which essentially serves as a feedback loop. This can be

mitigated by keeping the mosquitoes in many small groups (5 cups of 20 or even better 10

cups of 10, so an artifactual death spiral would only affect a single replicate cup) or by

removing dead individuals each day.

Authors:

Yes, we agree, and we were aware of this potential problem. The survival

experiments were conducted in five 500ml cups, each containing 20 mosquitoes. Dead

mosquitoes were removed on a daily basis.

New Comment: This is re-assuring, but this information is not included in the revised manuscript.

Reviewer #2:

The authors show that silencing of AaCBP is over by 10 DPI, and that levels of AMPs and

Vago2 are back to normal by 10 DPI (Fig S6). However, differences in ZIKV titer do not

emerge until 10 DPI (Fig 3), with no difference at 5 DPI-even though that is when silencing

of AaCBP and the AMPs are maximal. This is not consistent with AaCBP or the indicated

AMPs acting directly in anti-ZIKV immunity, and this inconsistency is not addressed in the

manuscript even though it goes to the heart of the main conclusions. Likewise in Figure 5,

NaB experiments show that AMPs are induced in ZIKV-infected individuals at 3 DPI with

reduced virus titer at 7 DPI, but not 15 DPI. Again, this is not consistent with the data

presented in Fig3, where the authors report an increase in viral titer at 15 DPI after

knockdown of CBP.

Authors:

The data of Figure S6 show the levels of mRNA transcripts, which are indeed

compromised until Day 5 and start to be restored at Day 10. However, it is difficult to

tell when proteins are fully expressed, translated and/or active to re-establish cellular

homeostasis. The same consideration can be assumed for the data of Figure 5.

In addition, please, see the data of Figure 2D and E. Note that in Aag2 cells, AaCBP

gene transcripts reach their peak at 6 hours post-infection, but the peak of its

enzymatic activity is only reached at 15 hours post-infection.

New Comment: A delay between RNA levels and protein levels may be able to explain a difference of hours, but not days. AMPs by their nature must be produced rapidly and then vanish when the threat is gone. No matter what the explanation, the manuscript must reflect these inconsistencies, rather than ignore them.

Reviewer #2:

In the latter, dsRNA experiments cannot disentangle the role of CBP in histone acetylation

from its more general role as CREB binding protein in RNA transcription.

Authors:

We are not aware of the function of CBP proteins (in all different model organisms thus far studied) outside its role as a transcription coactivator through its histone acetyltransferase activity. However, this is a clear possibility, and this possibility can certainly be addressed in our future studies.

New Comment: I am confused by this response, as the authors themselves describe CBPs as being “recruited to promoters by interaction with DNA-bound transcription factors, which directly interact with the RNA polymerase II complex.”. Thus, CBP is a physical scaffold protein, and the authors also state “The TAZ, KIX and CREB are the protein interaction domains that mediate interactions with transcription factors”. Thus, silencing of CBP depleted not just the HAT function, but also the scaffold function. This should be discussed.

New comment:

“These results indicate that even a partial deletion of AaCBP is enough to disrupt the homeostasis of the mosquito”- The phrase “partial deletion” here does not make sense, as nothing was deleted. Replace with “partial reduction”, “partial silencing” or similar phrase. Maybe the authors meant “partial depletion”?

--------------------

PLOS authors have the option to publish the peer review history of their article (what does this mean?). If published, this will include your full peer review and any attached files.

If you choose “no”, your identity will remain anonymous but your review may still be made public.

Do you want your identity to be public for this peer review? For information about this choice, including consent withdrawal, please see our Privacy Policy.

Reviewer #1: Yes: Sujit Pujhari

Reviewer #2: No

Figure Files:

While revising your submission, please upload your figure files to the Preflight Analysis and Conversion Engine (PACE) digital diagnostic tool, https://pacev2.apexcovantage.com. PACE helps ensure that figures meet PLOS requirements. To use PACE, you must first register as a user. Then, login and navigate to the UPLOAD tab, where you will find detailed instructions on how to use the tool. If you encounter any issues or have any questions when using PACE, please email us at figures@plos.org.

Data Requirements:

Please note that, as a condition of publication, PLOS' data policy requires that you make available all data used to draw the conclusions outlined in your manuscript. Data must be deposited in an appropriate repository, included within the body of the manuscript, or uploaded as supporting information. This includes all numerical values that were used to generate graphs, histograms etc.. For an example see here: http://www.plosbiology.org/article/info%3Adoi%2F10.1371%2Fjournal.pbio.1001908#s5.

Reproducibility:

To enhance the reproducibility of your results, we recommend that you deposit your laboratory protocols in protocols.io, where a protocol can be assigned its own identifier (DOI) such that it can be cited independently in the future. Additionally, PLOS ONE offers an option to publish peer-reviewed clinical study protocols. Read more information on sharing protocols at https://plos.org/protocols?utm_medium=editorial-email&utm_source=authorletters&utm_campaign=protocols

References

Please review your reference list to ensure that it is complete and correct. If you have cited papers that have been retracted, please include the rationale for doing so in the manuscript text, or remove these references and replace them with relevant current references. Any changes to the reference list should be mentioned in the rebuttal letter that accompanies your revised manuscript. If you need to cite a retracted article, indicate the article's retracted status in the References list and also include a citation and full reference for the retraction notice.

PLoS Negl Trop Dis. doi: 10.1371/journal.pntd.0010559.r005

Decision Letter 2

Philippe Desprès, Jason L Rasgon

3 Jun 2022

Dear Dr FANTAPPIE,

We are pleased to inform you that your manuscript 'Zika virus infection drives epigenetic modulation of immunity by the histone acetyltransferase CBP of Aedes aegypti' has been provisionally accepted for publication in PLOS Neglected Tropical Diseases.

Before your manuscript can be formally accepted you will need to complete some formatting changes, which you will receive in a follow up email. A member of our team will be in touch with a set of requests.

Please note that your manuscript will not be scheduled for publication until you have made the required changes, so a swift response is appreciated.

IMPORTANT: The editorial review process is now complete. PLOS will only permit corrections to spelling, formatting or significant scientific errors from this point onwards. Requests for major changes, or any which affect the scientific understanding of your work, will cause delays to the publication date of your manuscript.

Should you, your institution's press office or the journal office choose to press release your paper, you will automatically be opted out of early publication. We ask that you notify us now if you or your institution is planning to press release the article. All press must be co-ordinated with PLOS.

Thank you again for supporting Open Access publishing; we are looking forward to publishing your work in PLOS Neglected Tropical Diseases.

Best regards,

Jason L. Rasgon

Associate Editor

PLOS Neglected Tropical Diseases

Philippe Desprès

Deputy Editor

PLOS Neglected Tropical Diseases

***********************************************************

PLoS Negl Trop Dis. doi: 10.1371/journal.pntd.0010559.r006

Acceptance letter

Philippe Desprès, Jason L Rasgon

21 Jun 2022

Dear Dr. Prof. Fantappié,

We are delighted to inform you that your manuscript, "Zika virus infection drives epigenetic modulation of immunity by the histone acetyltransferase CBP of Aedes aegypti," has been formally accepted for publication in PLOS Neglected Tropical Diseases.

We have now passed your article onto the PLOS Production Department who will complete the rest of the publication process. All authors will receive a confirmation email upon publication.

The corresponding author will soon be receiving a typeset proof for review, to ensure errors have not been introduced during production. Please review the PDF proof of your manuscript carefully, as this is the last chance to correct any scientific or type-setting errors. Please note that major changes, or those which affect the scientific understanding of the work, will likely cause delays to the publication date of your manuscript. Note: Proofs for Front Matter articles (Editorial, Viewpoint, Symposium, Review, etc...) are generated on a different schedule and may not be made available as quickly.

Soon after your final files are uploaded, the early version of your manuscript will be published online unless you opted out of this process. The date of the early version will be your article's publication date. The final article will be published to the same URL, and all versions of the paper will be accessible to readers.

Thank you again for supporting open-access publishing; we are looking forward to publishing your work in PLOS Neglected Tropical Diseases.

Best regards,

Shaden Kamhawi

co-Editor-in-Chief

PLOS Neglected Tropical Diseases

Paul Brindley

co-Editor-in-Chief

PLOS Neglected Tropical Diseases

Associated Data

    This section collects any data citations, data availability statements, or supplementary materials included in this article.

    Supplementary Materials

    S1 Table. List of primers used in this study.

    (XLSX)

    S1 Fig. Protein domain and alignment of putative histone acetyltransferases from A. aegypti.

    A. Genome identification of putative lysine acetyltransferase (KAT) homologs from A. aegypti, AaHAT1 (XP_001651817.1), AaTip60 (XP_021706851.1), AaMOF1 (XP_001658578.2), AaMOF2 (XP_021711683.1), HBO1 (XP_021701314.1), GCN5 (XP_0016566424.1) and AaCBP itself. Functional domains are indicated above each box. B. Protein sequence alignment of the HAT domains from the putative A. aegypti KATs. Amino acids in red show identity or conservation among all 7 HAT domains.

    (TIF)

    S2 Fig. Blood meal upregulates the expression of AaCBP.

    Mosquitoes were fed with blood over different time courses and mRNA quantification by qPCR was performed in the midgut or fat body. The results in A and B are pools of at least 3 independent experiments, plotted using samples from sugar-fed mosquitoes as reference. Error bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05; **, p < 0.01; ***, p <0.001.

    (TIF)

    S3 Fig. Silencing levels of AaCBP after a sugar or blood meal.

    Two days after feeding, the AaCBP gene was knocked down and its expression was measured two days after silencing. qRT-PCR was performed from 3 independent biological replicates. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p < 0.05.

    (TIF)

    S4 Fig. Silencing of AaCBP in adult female mosquitoes by intrathoracic injections of dsRNAs.

    The expression (A) and activity (B) of AaCBP were used to evaluate the efficiency of gene knockdown. A. The mRNA levels of AaCBP in the midgut, ovary, fat body or whole mosquito were quantified by qRT-PCR at 48 h postinjection with dsCBP or dsLuc. Silencing level was determined by the ratio between mRNA levels of AaCBP-silenced versus dsLuc-injected mosquitoes. B. Western blot of 10 μg of total protein extract of dsCBP- or dsLuc-injected-mosquitoes. Monoclonal antibodies against acetylated- or nonacetylated histone H3 (loading control) are indicated. The intensity of the bands was quantified by densitometry analysis plotted as a graph using ImageJ (NIH Software). Western blotting was performed on 3 independent biological replicates and one representative is shown here. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05.

    (TIF)

    S5 Fig. ZIKV infection upregulates the expression of antimicrobial peptides in the midgut of A. aegypti.

    Fifty adult female mosquitoes were infected with ZIKV after feeding on infected blood and the expression of cecropin D (CEC D), cecropin G (CEC G), attacin (ATTC), defensin C (DEF C) and serpin (SERP) was measured by qRT-PCR four days post infection. qRT-PCR was performed from 3 independent biological replicates. Bars indicate the standard error of the mean.

    (TIF)

    S6 Fig. Silencing of AaCBP is sustainable and efficient until five days of ZIKV infection.

    Fifty adult female mosquitoes were infected with ZIKV for 1, 5, 7 or 10 days and the expression levels of AaCBP, cecropin G, defensin C or vago 2 in the fat body were measured by qRT-PCR on 3 independent biological replicates. Bars indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05.

    (TIF)

    S7 Fig. Intestinal microbiota load by genomic DNA qPCR.

    Fifty nanograms of genomic DNA from a pool of 15 midguts from silenced female mosquitoes was used as a template. Amplifications were carried out using specific primers for the ribosomal genes 16S and 18S, for bacteria or fungi, respectively. The A. aegypti ribosomal protein 49 gene (Rp49) was used as an endogenous control. Quantifications were carried out using the comparative Ct method. Genomic DNA qPCR was performed from 6 independent biological replicates. Bars indicate the standard error of the mean.

    (TIF)

    S8 Fig. Histone deacetylase inhibition leads to histone hyperacetylation in A. aegypti.

    Ten adult female mosquitoes were intrathoracically injected with PBS, 0.25 M (17.5 pmol), or 0.5 M (35 pmol) of sodium butyrate (NaB). Four days after treatment, 10 μg of total protein extract from 10 mosquitoes was used for histone acetylation analysis. Western blotting with monoclonal antibodies against H3K9ac, H3K27ac, H3 panacetylated, or H3 (as loading control) was performed. The intensity of the bands was quantified by densitometry (lower panel) analysis plotted as a graph using ImageJ (NIH Software). Western blotting was performed on 3 independent biological replicates and one representative is shown in panel A. Error bars in Panel B indicate the standard error of the mean; statistical analyses were performed by Student’s t test. *, p <0.05.

    (TIF)

    Attachment

    Submitted filename: Rebuttal letter PNTD-D-21-01702.pdf

    Attachment

    Submitted filename: Rebuttal letter PNTD #2.pdf

    Data Availability Statement

    All relevant data are within the manuscript and its Supporting Information files.


    Articles from PLoS Neglected Tropical Diseases are provided here courtesy of PLOS

    RESOURCES