Next Article in Journal
Mitochondrial Phylogenomics Suggests Complex Evolutionary Pattern of Pronotal Foliaceous Mimicry in Hierodulinae (Mantodea: Mantidae), with Description of a New Species of Rhombodera Burmeister, 1838 from China
Previous Article in Journal
The Functional Morphology of the Bursa Copulatrix of a Butterfly That Does Not Digest Spermatophores (Leptophobia aripa, Pieridae)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Transcriptome of Nosema ceranae and Upregulated Microsporidia Genes during Its Infection of Western Honey Bee (Apis mellifera)

1
Department of Entomology, National Chung Hsing University, Taichung City 40227, Taiwan
2
Department of Biotechnology and Animal Science, National Ilan University, Yi-Lan City 26047, Taiwan
3
Department of Computer Science and Engineering, Yuan-Ze University, Tao-Yuan City 32003, Taiwan
4
Agricultural Biotechnology Research Center, Academia Sinica, Taipei City 11529, Taiwan
5
Biotechnology Center in Southern Taiwan, Academia Sinica, Tainan 711010, Taiwan
*
Author to whom correspondence should be addressed.
Insects 2022, 13(8), 716; https://doi.org/10.3390/insects13080716
Submission received: 10 June 2022 / Revised: 29 July 2022 / Accepted: 5 August 2022 / Published: 9 August 2022
(This article belongs to the Topic Advanced in Honey Bee and Apitherapy)

Abstract

:

Simple Summary

In this study, the gene expression profile of a honey bee microsporidium, Nosema ceranae, was investigated at 5, 10, and 20 days post-infection. Based on transcriptome data, the common DEGs and stage-specific genes were identified and validated. Our data reveal that the gene expression of N. ceranae during infection is highly related to the mechanisms of gene transcription, protein synthesis, and structural proteins. This information could be a reference for the control of nosemosis at the genetic level in apiculture.

Abstract

Nosema ceranae is one of the fungal parasites of Apis mellifera. It causes physical and behavioral effects in honey bees. However, only a few studies have reported on gene expression profiling during A. mellifera infection. In this study, the transcriptome profile of mature spores at each time point of infection (5, 10, and 20 days post-infection, d.p.i.) were investigated. Based on the transcriptome and expression profile analysis, a total of 878, 952, and 981 differentially expressed genes (DEGs) (fold change ≥ 2 or ≤ −2) were identified in N. ceranae spores (NcSp) at 5 d.p.i., 10 d.p.i., and 20 d.p.i., respectively. Moreover, 70 upregulated genes and 340 downregulated genes among common DEGs (so-called common DEGs) and 166 stage-specific genes at each stage of infection were identified. The Gene Ontology (GO) analysis indicated that the DEGs and corresponding common DEGs are involved in the functions of cytosol (GO:0005829), cytoplasm (GO:0005737), and ATP binding (GO:0005524). Furthermore, the pathway analysis found that the DEGs and common DEGs are involved in metabolism, environmental information processing, and organismal systems. Four upregulated common DEGs with higher fold-change values, highly associated with spore proteins and transcription factors, were selected for validation. In addition, the stage-specific genes are highly involved in the mechanism of pre-mRNA splicing according to GO enrichment analysis; thus, three of them showed high expression at each d.p.i. and were also subjected to validation. The relative gene expression levels showed a similar tendency as the transcriptome predictions at different d.p.i., revealing that the gene expression of N. ceranae during infection may be related to the mechanism of gene transcription, protein synthesis, and structural proteins. Our data suggest that the gene expression profiling of N. ceranae at the transcriptomic level could be a reference for the monitoring of nosemosis at the genetic level.

1. Introduction

Microsporidia are considered primitive eukaryotic and obligate intracellular parasites; they have 70 S ribosomes of small subunit (SSU) and large subunit (LSU) rRNA, which is similar to eukaryotes, but they lack mitochondria, peroxisomes, and centrioles. Nonetheless, molecular phylogeny shows that microsporidia are regarded as highly specialized parasitic fungi [1,2,3,4,5,6]. Nosemosis is a serious and common microsporidial disease of adult honey bees that is mainly caused by two species of microsporidians, Nosema apis and Nosema ceranae [7,8]. N. ceranae and N. apis are described as midgut pathogens found in adult honey bees that reproduce only in midgut epithelial cells [9]. There are no clear outward symptoms of nosemosis; however, whitish and swollen midguts are observed in heavily infected honey bees, honey bees become weak, and colonies can even collapse.
In general, Nosema spp. have two distinct phases in their life cycle, including the proliferative phase (merogony) and sporogonic phase (sporogony). The morphology and life cycle of N. apis and N. ceranae are quite similar, and previous studies have also proven that N. apis and N. ceranae can infect Apis mellifera [9,10,11]. However, some different features are found between N. apis and N. ceranae, including genome size, temperature resilience, survival, and impacts on infected honey bees [9,12,13,14]. Compared with N. apis, N. ceranae has a higher virulence and appears to successfully compete with N. apis in many A. mellifera populations globally. Therefore, N. ceranae manifests as a long-term chronic infection in warmer areas and then causes colony loss [15,16,17].
As mentioned above, N. ceranae mainly infects the midgut tissues of honey bees, and many studies have demonstrated the impacts of an N. ceranae infection on A. mellifera. Once honey bees are infected with N. ceranae, immunity and physiological conditions decrease, eventually leading to colony weakness and losses [18,19]. Because of the pressure of the lack of nutrition and energy, a higher hunger level is observed in infected honey bees, and more sugar is consumed [19,20,21]. Moreover, N. ceranae replicates with considerable energy taken from the host, causing malnutrition and increased mortality in honey bees [22]. The expression profile of the genes associated with the carbohydrate metabolism of A. mellifera are changed during the process of an N. ceranae infection. Besides, severe physiological effects caused by an infection with N. ceranae, including the expression of age-inappropriate behaviors, a reduced flying capability, and a shortened lifespan, were also reported [23]. It was reported that an infection with N. ceranae does not cause large compositional changes to the honey bee midgut microbiota, but the relative abundances of Proteobacteria and Firmicutes were significantly changed under a severe N. ceranae infection, indicating the dysbiosis state of the bacteriome in the honey bee gut [24]. In addition, it has also been reported that co-exposure to pesticides, such as fipronil or thiamethoxam, and N. ceranae decreased the survival rate of honey bees [25]. In terms of microbiota composition, co-exposure to pesticides and N. ceranae also negatively influenced the abundance of Alphaproteobacteria, but increased the Gammaproteobacteria abundance, indicating that the pesticides and N. ceranae are associated with the pathobiome status of honey bees [25]. Even though some bacterial species exhibited inhibitory activity on N. ceranae, it has been reported that the enhancement of Bifidobacterium spp. in the hindgut after N. ceranae infection may not inhibit nosemosis [25,26,27].
Previous studies have examined the gene expression profile and corresponding differentially expressed genes of N. ceranae-infected A. mellifera [28,29,30]. However, few studies have emphasized either the comprehensive gene expression of N. ceranae during its infection of A. mellifera or the common DEGs, which are important to N. ceranae for completing its life cycle [31]. Moreover, the genome of N. ceranae has been studied and reported for a better understanding of the architecture, regulation, and evolution of this pathogen [31,32]. From our previous study, the identification of N. ceranae-specific genes during its infection of A. mellifera by suppressive subtractive hybridization (SSH) has been performed [33], but comprehensive data on the gene expression profile during the infection process are still unavailable. In this study, we attempted to reveal the transcriptomic profiling of N. ceranae during its infection of A. mellifera to complete the information on the infection cycle of this parasite. To this end, a transcriptomic approach at different N. ceranae infection stages in A. mellifera was followed, and common DEGs and stage-specific genes among different infectious periods were identified and validated. This study provides information to better understand the mechanism of N. ceranae and guide basic disease-control studies.

2. Materials and Methods

Honey bees
Worker honey bees were collected from A. mellifera colonies in apiaries at the National Chung Hsing University (NCHU, Taiching, Taiwan). Each honey bee colony had six to eight frames and the presence of a normal spawning queen. Brood cells of honeycombs from the six colonies, which were capped at more than 70–80%, were selected and incubated at 34 °C until honey bees emerged. The newly emerged adult honey bees were collected in a rearing cage and supplied with 50% syrup for the following experiments.
Artificial infection of Nosema ceranae
Mature N. ceranae spores were collected from the midgut tissues of honey bees from apiaries at the NCHU (National Chung Hsing University, Taichung, Taiwan) and purified by the protocol proposed by Huang et al., 2007 [34]. Briefly, the midgut tissues of honey bees were collected in 500 μL 1 × TE buffer (10 mM Tris, 1 mM EDTA, pH 7.5) (Bioman, New Taipei, Taiwan) and homogenized with a pestle. The homogenized tissues were then filtered with 100 mesh stainless steel wire cloth. The filtrate was centrifuged at 1000× g for 10 min, and the supernatant and tissue debris were removed. Then, the pellet was resuspended in 500 μL 1 × TE buffer. The wash step was repeated to remove any remaining tissue debris. The spore suspension was centrifuged at 16,000× g for 3 min at 4 °C, and the collected spores were counted with a haemocytometer (Paul Marienfeld GmbH & Co. KG, Baden-Württemberg, Germany). The concentration of the mature spore suspension was adjusted to 1 × 105 spore/2 μL for the infection of honey bees [35].
In the N. ceranae-infected group, the newly emerged honey bees were fed 2 μL of 50% sugar syrup with N. ceranae spores (1 × 105 spores per honey bee), while the honey bees in the control group were fed 2 μL of 50% sugar syrup. Each group contained fifty honey bees in one rearing cage, which were supplied with 50% syrup and incubated at 34 °C until sampling. Samples for both RNA-seq and the detection of N. ceranae infections were selected in triplicate, and qPCR validation for gene expression was repeated in four replicates. The survival rates of the infected and control groups were observed and recorded daily. A Kaplan–Meier survival curve analysis was performed using SPSS (Norman H. Nie, C. Hadlai (Tex) Hull and Dale H. Bent, IBM SPSS Statistics version 20, Palo Alto, CA, USA).
Confirmation of Nosema ceranae infection
Before the experiment, multiplex PCR [36] was performed to confirm the absence of Nosema spp. infection in newly emerged adult honey bees (Supplementary Figure S1). To confirm the N. ceranae infection, three midgut tissues of N. ceranae-infected bees and noninfected honey bees were collected in 1 × TE buffer. The midgut tissues were homogenized with a pestle, and the tissue suspension was subjected to observation, spore counting, and DNA extraction. Mature spores were observed under 400× light microscopy (WHITED, Taiwan) and counted by a haemocytometer (Paul Marienfeld GmbH & Co. KG, Baden-Württemberg, Germany). A total of 100 μL of homogenized tissue suspension was subjected to DNA extraction by the EDNA HISPEX (Easy DNA High-Speed Extraction) tissue kit (Fisher Biotec, Stirling, Australia) according to the user manual. The DNA sample was then subjected to PCR detection with the microsporidia universal primer set 18f/1537r, and the Hb18f/Hb18r primer set was used as an internal control (Supplementary Table S1). For the PCR, each 20 μL PCR contained 10 μL 2× SuperRed PCR Master mix (BIOTOOLs, New Taipei, Taiwan), 1 μL 100 mM of each primer, and 1 μL template DNA. PCR amplification was performed by a MiniAmp™ Thermal Cycler (Thermo Scientific, Waltham, MA, USA) with the following program: the thermal cycle was preheated to 95 °C for 10 min, followed by 35 cycles of 95 °C for 15 s, 55 °C for 30 s, and 72 °C for 1 min 30 s, and then a final extension was performed of 72 °C for 10 min and storage at 20 °C. The PCR product was analyzed by electrophoresis on a 1% agarose gel in 1 × Tris-acetate-EDTA (TAE) buffer.
Preparation of RNA samples for next-generation sequencing
Based on the data of the survival curve and mature spore counts, midgut tissues from three honey bees were collected at 0 (3 h), 5, 10, and 20 days post-infection (d.p.i.) and purified N. ceranae mature spores (1 × 108) were homogenized by a Tissue Lyser II (QIAGEN, Solingen, Germany) at f = 30/s for 1 min. This procedure was repeated at three time points. The total RNA was extracted using TRIzol reagent (Invitrogen Life Technologies, Waltham, MA, USA) following the user’s manual. The extracted RNA sample was then purified using a Locking Gel (BIOTOOLS, New Taipei, Taiwan). The RNA purity and integrity were checked by a Nanodrop 1000 Spectrophotometer V3.5 (Thermo Scientific, Waltham, MA, USA) and Qubit 2.0 Fluorometer (Invitrogen Life Technologies, Waltham, MA, USA), respectively. Total RNA samples of three biological repeats were pooled and used for library construction. The size distribution of each RNA sample was determined by Fragment Analyzer Systems (Agilent, Santa Clara, CA, USA). The RNA sample was then used to construct a library for transcriptome sequencing.
RNA-Seq library construction and sequencing
The mRNA from each sample was further purified for sequencing library construction using an Ultra RNA Library Prep Kit (NEB, Essex, MA, USA) according to the manufacturer’s protocol. Briefly, poly-A-containing mRNA was purified by oligo d (T) magnetic beads. During the final step of mRNA purification, RNA was fragmented and primed with NEBNext random primers for cDNA synthesis. After mRNA was reverse-transcribed, end-repair and adaptor ligation (NEBNext Adaptor) were performed. Finally, DNA fragments were selectively enriched by PCR and then qualified and quantified by Fragment Analyser Systems (Agilent, Santa Clara, CA, USA) and qPCR, respectively, for the DNA library quality control analysis. The libraries were then pooled and sequenced by the Illumina MiSeq sequencer with a paired-end (PE) technology of 301 bp in length to generate high-throughput transcriptome sequencing data. Sequencing adaptors were trimmed using Trimmomatic [37] from raw PE reads, and then a read quality check was performed by PRINseq [38] from trimmed PE reads. To confirm the pathogen load in the sequencing data of the newly emerged adult honey bees, the trimmed PE reads were mapped to the genomes of Nosema apis and 6 other viruses.
Differentially expressed gene (DEG) analysis
The raw PE reads were trimmed by Trimmomatic [37], and then the clean reads were obtained. The quality reads were mapped to the N. ceranae genome index (ASM98816v1) using HISAT2 with default parameters, and the corresponding gene annotation information (GCF_000988165.1_ASM98816v1_genomic.gff) from the National Biotechnology Information Center (NCBI) (https://www.ncbi.nlm.nih.gov/ (accessed on 19 January 2021)) was used to obtain the N. ceranae-related PE reads for further analysis. Differentially expressed genes (DEGs) among experimental groups were analyzed via Cuffdiff, and the gene expression heatmap was generated by the R package [39] (Supplementary Figure S2). For the DEG analysis, genes with a zero FPKM value from each group of the analysis were filtered out. To facilitate logarithmic transformation, a value of 1 was added to the FPKM value of the remaining genes. Raw signals (FPKM t 1) were then transformed using log2-based transformation. A statistical test of fold changes was performed, and a significant difference in the expression of these two groups was defined as a log2 ratio ≥ 1 or ≤ −1 (fold change ≥ 2 or ≤ −2). The DEGs from each comparison (NcSp/5 d.p.i., NcSp/10 d.p.i., and NcSp/20 d.p.i.) were further compared by Venny 2.1.0 (Juan Carlos Oliveros, Madrid, Spain) [40] to obtain the common DEGs of N. ceranae in different infection stages. The DEGs and common DEGs were further used in the GO enrichment analysis and submitted to the KEGG Automatic Annotation Server (KAAS, http://www.genome.jp/tools/kaas/ (accessed on 23 April 2021)) for gene orthologue assignment and pathway mapping [41]. The selected N. ceranae common DEGs were further analyzed by EMB BLAST (European Molecular Biology Network—Swiss node) for a functional homology search. The results with an E-value < 10−4 were recorded for further discussion of functional proteins.
qRT–PCR validation
For validation, three midgut tissues of infected honey bees were collected into 300 μL of TE buffer at 0.5 (12 h), 1, 3, 5, 10, 15, and 20 d.p.i., with four replicates, and were homogenized for RNA extraction. Four replicates of 1 × 108 purified mature N. ceranae spores were also subjected to RNA extraction. The total RNA for qRT–PCR validation was extracted from 200 μL of homogenized midgut tissue and the N. ceranae mature spore sample using the GENEzol™ TriRNA Pure Kit (Geneaid, New Taipei, Taiwan) as described in the user manual. The concentration and quantity of the total RNA were measured using a Nanodrop™ spectrophotometer (Thermo Scientific, Waltham, USA). Each RNA sample was adjusted to 1 µg/mL for cDNA synthesis by BioDiamond RT–PCR mix (Origin Pure BioSci & Tech.Co., Yilan, Taiwan) as described in the user manual. Four highly upregulated common DEGs, AAJ76_1600052943, AAJ76_2000149161, AAJ76_2000141845, and AAJ76_500091146, and three highly expressed stage-specific genes, AAJ76_5000026485, AAJ76_2400020190, and AAJ76_2300035467, were selected for RT–qPCR validation. β-Tubulin from N. ceranae served as the internal control. Real-time PCR was performed by iQ™ SYBR® Green Supermix (Bio–Rad, Hercules, CA, USA) in a CFX Connect Real-Time PCR Detection System (Bio–Rad, Hercules, CA, USA). qRT–PCR was performed as follows: 95 °C for 3 min, followed by 40 cycles of 95 °C for 10 s, 60 °C for 30 s, and 72 °C for 45 s. The relative gene expression levels for common DEGs were calculated by the 2-△△Ct method, and stage-specific genes were calculated by the 2-△Ct method [42].
Statistical analysis
Statistical analysis and plotting were performed with R 4.0.3 [43] and the heatmap, ggplot2, and ggpubr [39,44,45] packages. The statistical analysis of the Kaplan–Meier survival curves was performed using the R packages survminer [46] and survival [47]. The statistical analysis of the spore count and qRT–PCR data was performed using one-way ANOVA with SPSS® [48].

3. Results

3.1. N. Ceranae Infection

The Kaplan–Meier survival curve results showed that survival fitness was significantly reduced (p < 0.01) in the N. ceranae-infected A. mellifera (Figure 1A). Mature spores were observed from 5 to 20 d.p.i. in the homogenized midgut tissues of the N. ceranae-infected honey bees under microscopy; while only a few mature spores were observed at 5 d.p.i., the spore counts significantly increased from 10 to 20 d.p.i (Figure 1B,C). Molecular detection was performed to detect the infection process of N. ceranae on A. mellifera. Although the earliest time that mature spores were observed was at 5 d.p.i., the results showed that N. ceranae infection was first detected at 3 d.p.i. by PCR (Figure 1D).

3.2. Sequencing Data Summary

The sequencing data are summarized in Supplementary Table S2. A total of 1.68 million (M), 6.85 M, 1.79 M, 1.58 M, and 2.91 M raw PE reads were obtained from 0 (3 h), 5, 10, and 20 d.p.i. and mature N. ceranae spores. The number of clean reads was 1.44 M, 2.35 M, 1.57 M, 1.38 M, and 1.58 M for 0 (3 h), 5, 10, and 20 d.p.i. and mature N. ceranae spores, respectively. The mappability of each sample to the N. ceranae genome was 0%, 4%, 3%, 3%, and 12% for 0 (3 h), 5, 10, and 20 d.p.i. and mature N. ceranae spores, respectively. In addition, the newly emerged adult honey bees were confirmed to be free of N. apis and common honey bee viral infections (Supplementary Table S3). These mapped reads were subjected to further analysis of the expression of different genes.

3.3. DEG Analysis

For the DEG analysis, a total of 878, 952, and 981 genes were identified to have significant fold changes (fold change ≥ 2 or ≤−2) in NcSp/5 d.p.i., NcSp/10 d.p.i., and NcSp/20 d.p.i., respectively. Among these DEGs, 429, 117, and 113 genes were upregulated and 449, 835, and 868 genes were downregulated in NcSp/5 d.p.i., NcSp/10 d.p.i., and NcSp/20 d.p.i., respectively (Figure 2A; Table 1, Table 2, and Tables S4–S9). The common DEGs were selected from all three groups of common DEGs by set intersection, and stage-specific genes were defined as the genes that were only expressed at 5, 10, and 20 d.p.i., respectively (Table 3 and Table S10). There was a total of 70 upregulated and 340 downregulated common DEGs (Figure 2B; Table 1, Table 2, and Tables S4–S9), and the fold-change distribution at each infection stage is shown in Figure 2B. The results also showed that a greater upregulation of N. ceranae common DEGs and stage-specific genes was found in the initial infectious stage, whereas more genes of N. ceranae were suppressed in the late stage of infection (Figure 2). For the stage-specific genes, a total of 25, 71, and 70 genes were identified at each time post-infection (Supplementary Table S10), and the top 10 expression values (FPKM) of the stage-specific genes are summarized in Table 3.

3.4. Gene Ontology (GO) Enrichment Analysis

The results of the GO annotation are shown in Figure 3. At NcSp/5 d.p.i., some genes belonging to GO pathways were significantly upregulated, including cytosol (GO:0005829), cytoplasm (GO:0005737), ATP binding (GO:0005524), protein binding (GO:0005515), and translation (GO:0006412), which were commonly found in NcSp/5 d.p.i., NcSp/10 d.p.i., and NcSp/20 d.p.i.; meanwhile, the genes involved in the aforementioned GO terms were reduced in number and downregulated in the NcSp/10 d.p.i. and NcSp/20 d.p.i. groups (Figure 3 and Supplementary Tables S11–S16). For the GO annotation of common DEGs, cytosol (GO:0005829) and cytoplasm (GO:0005737) were also found in the upregulated common DEGs group; however, metal ion binding (GO:0046872), GTP binding (GO:0005525), and the ubiquitin-dependent protein catabolic process (GO:0006511) showed high gene counts as observed in the downregulated common DEGs (Figure 4A and Supplementary Tables S11–S16). On the other hand, many stage-specific genes of N. ceranae were involved in pre-mRNA-associated pathways, including “RNA binding” (GO:0003723) and “mRNA splicing, via spliceosome” (GO:0000398) (Figure 4B).

3.5. KEGG Analysis

The results of the KEGG analysis show the pathway annotation of the DEGs and common DEGs at different stages and indicate the mechanisms of N. ceranae infection (Figure 5 and Supplementary Tables S17–S22). In the comparison of NcSp at 5 d.p.i., there were 211 pathways that included upregulated genes and 200 pathways that included downregulated genes (Figure 5A). Among these pathways, highly abundant N. ceranae DEGs were involved in the metabolism (KO:09100), environmental information processing (KO:09130), and organismal systems (KO:09150) categories, including the carbohydrate metabolism (KO:09101), signal transduction (KO:09132), and immune system (KO:09151) pathways (Figure 5A and Supplementary Tables S17 and S18). In the NcSp at 10 d.p.i. group, 112 pathways included upregulated genes and 1224 pathways included downregulated genes. Among these pathways, three categories, including the metabolism pathway (KO:09100), human diseases (KO:09160), and environmental information processing (KO:09130) were the most mapped, and they contained the pathways of carbohydrate metabolism (KO:09101) and signal transduction (KO:09132) (Figure 5B; Supplementary Tables S19 and S20). For the NcSp/20 d.p.i. group, there were 129 pathways that included upregulated genes and 267 pathways that included downregulated genes, and the pathway categories that highly regulated genes belonged to, including metabolism pathways (KO:09100), environmental information processing (KO:09130), human diseases (KO:09160), and organismal systems (KO:09150), were identified. These pathways also included carbohydrate metabolism (KO:09101) and signal transduction (KO:09132) (Figure 5C; Supplementary Tables S21 and S22).
The KEGG analysis results of the common DEGs were similar to those of the DEGs (Figure 5D; Supplementary Tables S23 and S24). A total of 80 pathways included upregulated genes and 199 pathways included downregulated genes during N. ceranae infection (Figure 5D). A high number of downregulated common DEGs were involved in the pathways of signal transduction (KO:09132); protein families: genetic information processing (KO:09182) and carbohydrate metabolism (KO:09101) within the environmental information process (KO:09130); brite hierarchies (KO:09180); and metabolism (KO:09100) categories.

3.6. Validation

Four highly upregulated common DEGs (AAJ76_1600052943, AAJ76_2000149161, AAJ76_2000141845, and AAJ76_500091146) and three highly expressed stage-specific genes (AAJ76_5000026485, AAJ76_2400020190, and AAJ76_2300035467) at each time post-infection were selected to validate the RNA-seq results during infection by N. ceranae. The results showed that all of the upregulated common DEGs increased at 3 d.p.i. and reached a high peak at 15 to 20 d.p.i., while no or a low abundance of gene expression was detected at 0.5 to 1 d.p.i. (Figure 6). The validation result matched the detection of N. ceranae infection, which was detectable at 3 d.p.i. (Figure 1D), suggesting that N. ceranae begins the infection process and initiates gene expression at 3 d.p.i. In addition, the upregulated common DEGs, reaching a high peak at 15–20 d.p.i., also showed coordination with the increasing spore count of N. ceranae (Figure 1C).
For the stage-specific gene validations, the stage-specific genes were only expressed at specific times post-infection, while the gene expression of AAJ76_5000026485, AAJ76_2400020190, and AAJ76_2300035467 was detected at all times post-infection, including in mature spores (Figure 7). However, the results of three stage-specific genes showed higher expression levels at 5, 10, and 20 d.p.i., revealing a similar expression trend as the transcriptomic prediction (Figure 7).

4. Discussion

The pattern of N. ceranae infection in this study was similar to a previous report that N. ceranae was detectable in an early infection stage at 3 d.p.i., while mature spores were only observed at 5 d.p.i. [18,49]. Negative impacts on the survival fitness of N. ceranae-infected honey bees have been reported, and they explain the changes to immune suppression, nutritional deficiencies, and hormone regulation [18,50,51,52]. These data indicate that infected N. ceranae might begin to affect and have consistent negative effects on honey bees at earlier infectious stages. Altogether, these data reveal that N. ceranae infections can be detected as early as 3 d.p.i. and the infection level dramatically increases from 10 to 20 d.p.i.
The gene expression profiling of N. ceranae was described at the transcriptomic level in this study. The results indicated that a greater upregulation of N. ceranae’s common and stage-specific genes was found in the initial infectious stage, whereas more genes of N. ceranae were suppressed in the late stage of infection. This tendency was similar to the study of N. ceranae gene expression in Apis cerana, suggesting that the N. ceranae genes were expressed at an early infection stage, followed by the gene expression profile of N. ceranae switching from the merogony to sporogony stages [53]. Microsporidia lack mitochondria, and the loss of most metabolic pathways might involve an alteration of gene expression during the course of infection; therefore, host-generated ATP would be imported to support their metabolism [54,55]. Several highly expressed stage-specific genes (AAJ76_5000026485, AAJ76_2400020190, AAJ76_2300035467, and AAJ76_600074061) were identified as pre-mRNA splicing factors, indicating that post-transcriptional regulation was continually activated during N. ceranae infection.
For the GO annotation of common DEGs, cytosol (GO:0005829) and cytoplasm (GO:0005737) were also found in the upregulated common DEG group. The cytoplasm mainly comprises microsporidia, containing many ribosomes in a regular array [56]. During the replication of sporonts, division by binary fission produces two sporoblasts and infects the hosts through karyokinesis and cytokinesis [2]. Therefore, the upregulation of genes belonging to the cytoplasm with corresponding GO terms might contribute to the early propagation of N. ceranae. At the initiation of microsporidial infection, sporoplasm was injected into the cytoplasm of the host cell by the germination of the polar tube [2]. Because of the lack of genes for many metabolic pathways in the microsporidia, the microsporidia acquire nutrients from the host. Microsporidia in contact with the host cytoplasm could provide more opportunities to export cytotoxic compounds across the plasma membrane as an alternative to compensate for development [57]. In addition, upregulated genes belonging to the ATP-binding (GO:0005524) pathway were found in the early stage of N. ceranae infection (NcSp/5 d.p.i.). In the ATP-binding (GO:0005524) pathway, the transmembrane protein superfamily of microsporidia, which includes ATP-binding cassette (ABC) transporters, plays an important role in substrate transportation and pathogen development, indicating that energy transport is abundant in this stage [57].
It was noted that metal ion binding (GO:0046872), GTP binding (GO:0005525), and the ubiquitin-dependent protein catabolic process (GO:0006511) showed high gene counts as observed in the downregulated common DEGs. Metal ions are vital elements for proper folding and the ribozyme catalysis of nucleic acids [58,59,60]. GTP-binding proteins regulate cellular processes for multiple protein biosynthesis and intracellular membrane traffic [61]. The role of the ubiquitin–proteasome pathway has been reported to be the selective degradation of soluble cellular proteins under most conditions [62]. The downregulation of the genes belonging to the GO term indicate the requirement of nucleotide and protein bases from host cells in N. ceranae infections. In contrast, many stage-specific genes of N. ceranae were involved in pre-mRNA-associated pathways, including “RNA binding” (GO:0003723) and “mRNA splicing, via spliceosome” (GO:0000398), suggesting the impacts on the transcriptional level for N. ceranae infection.
Based on the result of KEGG pathway analysis on DEGs, the categories of metabolism and environmental information processing included a high number of downregulated genes. Moreover, in the pathways of metabolism, the “carbohydrate metabolism” included a high number of downregulated genes at each infection stage. It has been reported that Nosema acquires carbohydrates from the epithelial cells of the honey bee gut lining, leading to a series of metabolic changes in the host [3,20]; therefore, this point was also proven by the downregulation of genes belonging to “carbohydrate metabolism”.
Instead of mitochondria, microsporidia contain a genome-less organelle called a mitosome, which encodes only 20 mitosomal proteins in microsporidia, indicating the loss of function of the mitosome organ compared to that of normal mitochondria in the yeast Saccharomyces cerevisiae [32,63]; therefore, the mature spores could still maintain basic metabolism by their own system. In the beginning of infection, genes belonging to the “folding, sorting, and degradation (K09123)” and “replication and repair (K09124)” pathways in the genetic information processing (K09120) categories, the cellular processes pathway (KO:09140) of “cell growth and death” (K09143), and the environmental information processing pathway (KO:09130) of “signal transduction” (K09132) were highly upregulated in the NcSp/5 d.p.i. group. Meanwhile, the genes involved in these pathways were downregulated at a late stage of N. ceranae infection (NcSp/10 d.p.i. and NcSp/20 d.p.i. groups), suggesting that the merogony stage needs to initiate more functions of genetic and information processes than mature spores to facilitate the propagation of N. ceranae and maintain the activities of infection. Lastly, from our data, the downregulated genes involved in the pathway of carbohydrate metabolism suggest that N. ceranae takes energy from honey bees during infection [8].
The validation of four highly upregulated common DEGs and three highly expressed stage-specific genes at each time post-infection showed similar tendencies to those of the DEG predictions. The AAJ76_1600052943 and AAJ76_2000141845 showed a functional homology to the spore wall proteins SWP25 and SWP30, respectively. In Nosema bombycis, the two spore wall proteins have been characterized as an endosporal protein (SWP30) and an exosporal protein (SWP32) [64,65]. These two spore wall proteins bind to the chitin component of N. bombycis, implying that these proteins are the structural support proteins in the spore wall. SWP25 was first found in N. bombycis; the localization of SWP25 is the endospore, and this protein possesses a signal peptide and a heparin-binding motif. In addition, SWP25 is critical for host-to-host transmission and persistence in the environment [66]. Furthermore, SWP25 has been reported to be related to spore wall integrity and germination, and recent research has shown that the induction of SWP25 expression after binding with haematopoietin may change the integrity of pathogens and affect the clearance of immune cells or the transmission of pathogens in haemolymph [65,67]. Therefore, the upregulation of AAJ76_1600052943 and AAJ76_2000141845 might also play roles similar to those mentioned above to facilitate spore germination and mature spore development. The highly expressed SPW was consistent with our previous study on N. ceranae gene expression in A. mellifera [33] and the research on the N. ceranae gene expression in A. cerana, suggesting a significant role for SPW during N. ceranae infection [53].
AAJ76_2000149161 was mapped to a DNA-dependent protein kinase catalytic subunit (DNA-PKcs) by SWISS-MODEL [68]. DNA-PKcs is a key member of the phosphatidylinositol-3 kinase-like (PIKK) family of protein kinases and plays an essential role in the repair of DNA double-strand breaks by nonhomologous end joining (NHEJ) [69]. The upregulation of this gene might contribute to the relative function of N. ceranae in using host DNA fragments for its replication. AAJ76_500091146 is described as a forkhead hnf3 transcription factor based on protein functional prediction. HNF-3/FKH genes are a large family of transcriptional activators, and all forkhead proteins contain a highly conserved DNA binding domain and play numerous regulatory roles in eukaryotes. At least four genes in yeast are involved in a wide range of biological functions, such as apoptosis, cell cycle, immunity, development, growth, stress resistance, metabolism, and reproduction [70,71]. Based on the above functions, the upregulated AAJ76_500091146 might be involved in multiple regulatory roles in the development and growth of N. ceranae.
Interestingly, the functional prediction of all three stage-specific genes indicated post-transcriptional modification. AAJ76_5000026485 is a functional homologue of the pre-mRNA-splicing factor cwf14, which is a type of spliceosome that catalyzes the excision of introns from pre-mRNA [72]. A study from yeast showed that cwf14 functions in the regulation of cell growth and mRNA splicing and is also a key factor regulating cell wall biogenesis [73]. The function of AAJ76_2400020190 was found to match that of the putative H/ACA ribonucleoprotein complex subunit 3, which is known as a small nuclear RNP (snRNP). snRNP predominantly guides the site-directed pseudouridylation of target RNAs, processes ribosomal RNA, and stabilizes telomerase RNA [74]. Moreover, the function of H/ACA ribonucleoproteins is essential for ribosome biogenesis, pre-mRNA splicing, and telomere maintenance [74]. AAJ76_2300035467 is functionally similar to U6 snRNA-associated Sm-like protein LSm6. The Sm proteins are RNA–protein complex proteins, and LSm6 is an Sm-like protein (LSM) in S. cerevisiae. A study of S. cerevisiae suggested that LSm6 is associated with U6 snRNA and is required for the maintenance of normal U6 snRNA levels and for pre-mRNA splicing [75]. In summary, the primary transcription products (pre-mRNAs) undergo a variety of post-transcriptional modifications to create functional mRNAs [76]. These three stage-specific genes were identified at different infection stages, and all associations with pre-mRNA splicing suggest that they are highly involved in the activation of post-transcriptional modifications of N. ceranae at each infectious stage.

5. Conclusions

In this study, the transcriptomic analysis results indicated alterations in the N. ceranae gene expression status during A. mellifera infection. The identification of the common and stage-specific genes expressed during the infection process revealed the interaction between N. ceranae and A. mellifera. Downregulated genes during the infection, especially at a late period of infection, were found. Regarding RT–qPCR validation, the common DEGs showed higher expression levels from 5 d.p.i. to 20 d.p.i., indicating the importance of common DEGs. In our study, we attempted to identify the DEGs resulting from gene expression profiling based on transcriptomic analysis and proposed the common and stage-specific genes of N. ceranae during its infection of A. mellifera. Our data also describe the relatively comprehensive gene expression of the N. ceranae infection stage and analyze the details during infection, thus providing basic information on the N. ceranae mechanism during infection and furthering the application, including the early detection of nosemosis.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/insects13080716/s1, Figures S1 and S2; Tables S1–S24. Figure S1. The detection of Nosema spp. of newly emerged adult honey bees by multiplex PCR before experiment. The information of primer sets were listed in Supplementary Table S1; Figure S2. The pipeline of transcriptome analysis. The commend lines and analysis purposes are noted pairwise the item. DEG = differential expressed gene; FPKM = fragment per kilobase per million; GO = gene ontology; KEGG = Kyoto Encyclopedia of Genes and Genomes; Table S1. Primer sets list; Table S2. RNA sequencing summary; Table S3. Transcriptomic data for honey bee infected with Nosema ceranae were screened by mapping to the genomes of Nosema apis and 6 honey bee infecting viruses; Table S4. Up-regulated DEG list of NcSp/5 d.p.i. comparison group; Table S5. Down-regulated DEG list of NcSp/5 d.p.i. comparison group; Table S6. Up-regulated DEG list of NcSp/10 d.p.i. comparison group; Table S7. Down-regulated DEG list of NcSp/10 d.p.i. comparison group; Table S8. Up-regulated DEG list of NcSp/20 d.p.i. comparison group; Table S9. Down-regulated DEG list of NcSp/20 d.p.i. comparison group; Table S10. Non-core genes list at each time post infection; Table S11. Up-regulated GO list of NcSp/5 d.p.i. comparison group; Table S12. Down-regulated GO list of NcSp/5 d.p.i. comparison group; Table S13. Up-regulated GO list of NcSp/10 d.p.i. comparison group; Table S14. Down-regulated GO list of NcSp/10 d.p.i. comparison group; Table S15. Up-regulated GO list of NcSp/20 d.p.i. comparison group; Table S16. Down-regulated GO list of NcSp/20 d.p.i. comparison group; Table S17. Up-regulated KEGG list of NcSp/5 d.p.i. comparison group; Table S18. N.c spore vs. 5 d.p.i. Down regulated KEGG list; Table S19. N.c spore vs. 10 d.p.i. Up regulated KEGG list; Table S20. N.c spore vs. 10 d.p.i. Down regulated KEGG list; Table S21. N.c spore vs. 20 d.p.i. Up regulated KEGG list; Table S22. N.c spore vs. 20 d.p.i. Down regulated KEGG list; Table S23. Up-regulated KEGG list of core genes; Table S24. Down-regulated KEGG list of core genes.

Author Contributions

Conceptualization, Y.-H.L. and Y.-S.N.; methodology, Y.-H.L., Z.-T.C., T.-H.C. and J.-C.C.; software, Y.-H.L., Y.-F.H. and M.-R.Y.; validation, Y.-H.L.; formal analysis, Y.-H.L. and M.-R.Y.; investigation, Y.-H.L.; resources, M.-C.W. and Y.-L.Y.; data curation, Y.-H.L.; writing—original draft preparation, Y.-H.L.; writing—review and editing, Y.-F.H., M.-C.W., Y.-W.C. and Y.-S.N.; supervision, Y.-S.N.; project administration, Y.-S.N.; funding acquisition, Y.-S.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by 109-2313-B-005-048-MY3 from the Ministry of Science and Technology (MOST).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The RNA-Seq data were deposited in NCBI with BioProject ID: PRJNA814859.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Keeling, P. Five questions about microsporidia. PLoS Pathog. 2009, 5, e1000489. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Bigliardi, E.; Sacchi, L. Cell biology and invasion of the microsporidia. Microbes Infect. 2001, 3, 373–379. [Google Scholar] [CrossRef]
  3. Keeling, P.J.; Fast, N.M. Microsporidia: Biology and evolution of highly reduced intracellular parasites. Annu. Rev. Microbiol. 2002, 56, 93–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Corradi, N.; Keeling, P.J. Microsporidia: A journey through radical taxonomical revisions. Fungal Biol. Rev. 2009, 23, 1–8. [Google Scholar] [CrossRef]
  5. Canning, E.U. Nuclear division and chromosome cycle in microsporidia. Biosystems 1988, 21, 333–340. [Google Scholar] [CrossRef]
  6. Fries, I.; Martin, R.; Meana, A.; García-Palencia, P.; Higes, M. Natural infections of Nosema ceranae in European honey bees. J. Apic. Res. 2006, 45, 230–233. [Google Scholar] [CrossRef]
  7. Chen, Y.; Evans, J.D.; Zhou, L.; Boncristiani, H.; Kimura, K.; Xiao, T.; Litkowski, A.; Pettis, J.S. Asymmetrical coexistence of Nosema ceranae and Nosema apis in honey bees. J. Invertebr. Pathol. 2009, 101, 204–209. [Google Scholar] [CrossRef]
  8. Martín-Hernández, R.; Bartolomé, C.; Chejanovsky, N.; Le Conte, Y.; Dalmon, A.; Dussaubat, C.; García-Palencia, P.; Meana, A.; Pinto, M.A.; Soroker, V. Nosema ceranae in Apis mellifera: A 12 years postdetection perspective. Environ. Microbiol. 2018, 20, 1302–1329. [Google Scholar] [CrossRef] [Green Version]
  9. Huang, W.-F.; Solter, L.F. Comparative development and tissue tropism of Nosema apis and Nosema ceranae. J. Invertebr. Pathol. 2013, 113, 35–41. [Google Scholar] [CrossRef]
  10. Antúnez, K.; Martín-Hernández, R.; Prieto, L.; Meana, A.; Zunino, P.; Higes, M. Immune suppression in the honey bee (Apis mellifera) following infection by Nosema ceranae (Microsporidia). Environ. Microbiol. 2009, 11, 2284–2290. [Google Scholar] [CrossRef]
  11. Chen, Y.-W.; Chung, W.-P.; Wang, C.-H.; Solter, L.F.; Huang, W.-F. Nosema ceranae infection intensity highly correlates with temperature. J. Invertebr. Pathol. 2012, 111, 264–267. [Google Scholar] [CrossRef]
  12. Higes, M.; Martín, R.; Meana, A. Nosema ceranae, a new microsporidian parasite in honeybees in Europe. J. Invertebr. Pathol. 2006, 92, 93–95. [Google Scholar] [CrossRef]
  13. Williams, G.R.; Shutler, D.; Burgher-MacLellan, K.L.; Rogers, R.E. Infra-population and-community dynamics of the parasites Nosema apis and Nosema ceranae, and consequences for honey bee (Apis mellifera) hosts. PLoS ONE 2014, 9, e99465. [Google Scholar] [CrossRef] [Green Version]
  14. Fries, I. Nosema ceranae in European honey bees (Apis mellifera). J. Invertebr. Pathol. 2010, 103, S73–S79. [Google Scholar] [CrossRef]
  15. Perry, C.J.; Søvik, E.; Myerscough, M.R.; Barron, A.B. Rapid behavioral maturation accelerates failure of stressed honey bee colonies. Proc. Natl. Acad. Sci. USA 2015, 112, 3427–3432. [Google Scholar] [CrossRef] [Green Version]
  16. Blažytė-Čereškienė, L.; Arbačiauskienė, V.S.; Radžiutė, S.; Nedveckytė, I.; Būda, V. Honey bee infection caused by Nosema spp. in Lithuania. J. Apic. Sci. 2016, 60, 77–88. [Google Scholar] [CrossRef] [Green Version]
  17. Ilyasov, R.; Gaifullina, L.; Saltykova, E.; Poskryakov, A.; Nikolaenko, A. Defensins in the honeybee antiinfectious protection. J. Evol. Biochem. Physiol. 2013, 49, 1–9. [Google Scholar] [CrossRef]
  18. Higes, M.; García-Palencia, P.; Martín-Hernández, R.; Meana, A. Experimental infection of Apis mellifera honeybees with Nosema ceranae (Microsporidia). J. Invertebr. Pathol. 2007, 94, 211–217. [Google Scholar] [CrossRef]
  19. Martín-Hernández, R.; Botías, C.; Barrios, L.; Martínez-Salvador, A.; Meana, A.; Mayack, C.; Higes, M. Comparison of the energetic stress associated with experimental Nosema ceranae and Nosema apis infection of honeybees (Apis mellifera). Parasitol. Res. 2011, 109, 605–612. [Google Scholar] [CrossRef]
  20. Mayack, C.; Naug, D. Energetic stress in the honeybee Apis mellifera from Nosema ceranae infection. J. Invertebr. Pathol. 2009, 100, 185–188. [Google Scholar] [CrossRef]
  21. Alaux, C.; Brunet, J.L.; Dussaubat, C.; Mondet, F.; Tchamitchan, S.; Cousin, M.; Brillard, J.; Baldy, A.; Belzunces, L.P.; Le Conte, Y. Interactions between Nosema microspores and a neonicotinoid weaken honeybees (Apis mellifera). Environ. Microbiol. 2010, 12, 774–782. [Google Scholar] [CrossRef] [Green Version]
  22. Mayack, C.; Naug, D. Individual energetic state can prevail over social regulation of foraging in honeybees. Behav. Ecol. Sociobiol. 2013, 67, 929–936. [Google Scholar] [CrossRef]
  23. Goblirsch, M. Nosema ceranae disease of the honey bee (Apis mellifera). Apidologie 2018, 49, 131–150. [Google Scholar] [CrossRef] [Green Version]
  24. Jabal-Uriel, C.; Alba, C.; Higes, M.; Rodríguez, J.M.; Martín-Hernández, R. Effect of Nosema ceranae infection and season on the gut bacteriome composition of the European honeybee (Apis mellifera). Sci. Rep. 2022, 12, 9326. [Google Scholar] [CrossRef]
  25. Paris, L.; Peghaire, E.; Mone, A.; Diogon, M.; Debroas, D.; Delbac, F.; El Alaoui, H. Honeybee gut microbiota dysbiosis in pesticide/parasite co-exposures is mainly induced by Nosema ceranae. J. Invertebr. Pathol. 2020, 172, 107348. [Google Scholar] [CrossRef]
  26. Zhang, Y.; Lu, X.; Huang, S.; Zhang, L.; Su, S.; Huang, W.-F. Nosema ceranae infection enhances Bifidobacterium spp. abundances in the honey bee hindgut. Apidologie 2019, 50, 353–362. [Google Scholar] [CrossRef]
  27. Rubanov, A.; Russell, K.A.; Rothman, J.A.; Nieh, J.C.; McFrederick, Q.S. Intensity of Nosema ceranae infection is associated with specific honey bee gut bacteria and weakly associated with gut microbiome structure. Sci. Rep. 2019, 9, 3820. [Google Scholar] [CrossRef]
  28. Chaimanee, V.; Chantawannakul, P.; Chen, Y.; Evans, J.D.; Pettis, J.S. Differential expression of immune genes of adult honey bee (Apis mellifera) after inoculated by Nosema ceranae. J. Insect Physiol. 2012, 58, 1090–1095. [Google Scholar] [CrossRef]
  29. Aufauvre, J.; Misme-Aucouturier, B.; Viguès, B.; Texier, C.; Delbac, F.; Blot, N. Transcriptome analyses of the honeybee response to Nosema ceranae and insecticides. PLoS ONE 2014, 9, e91686. [Google Scholar] [CrossRef]
  30. Badaoui, B.; Fougeroux, A.; Petit, F.; Anselmo, A.; Gorni, C.; Cucurachi, M.; Cersini, A.; Granato, A.; Cardeti, G.; Formato, G.; et al. RNA-sequence analysis of gene expression from honeybees (Apis mellifera) infected with Nosema ceranae. PLoS ONE 2017, 12, e0173438. [Google Scholar] [CrossRef]
  31. Huang, Q.; Wu, Z.H.; Li, W.F.; Guo, R.; Xu, J.S.; Dang, X.Q.; Ma, Z.G.; Chen, Y.P.; Evans, J.D. Genome and evolutionary analysis of Nosema ceranae: A microsporidian parasite of honey bees. Front. Microbiol. 2021, 12, 1303. [Google Scholar] [CrossRef] [PubMed]
  32. Cornman, R.S.; Chen, Y.P.; Schatz, M.C.; Street, C.; Zhao, Y.; Desany, B.; Egholm, M.; Hutchison, S.; Pettis, J.S.; Lipkin, W.I.; et al. Genomic analyses of the microsporidian Nosema ceranae, an emergent pathogen of honey bees. PLoS Pathog. 2009, 5, e1000466. [Google Scholar] [CrossRef] [PubMed]
  33. Chang, Z.T.; Ko, C.Y.; Yen, M.R.; Chen, Y.W.; Nai, Y.S. Screening of Differentially Expressed Microsporidia Genes from Nosema ceranae Infected Honey Bees by Suppression Subtractive Hybridization. Insects 2020, 11, 199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Huang, W.-F.; Jiang, J.-H.; Chen, Y.-W.; Wang, C.-H. A Nosema ceranae isolate from the honeybee Apis mellifera. Apidologie 2007, 38, 30–37. [Google Scholar] [CrossRef]
  35. Huang, W.-F.; Solter, L.F.; Yau, P.M.; Imai, B.S. Nosema ceranae escapes fumagillin control in honey bees. PLoS Pathog. 2013, 9, e1003185. [Google Scholar] [CrossRef] [Green Version]
  36. Fries, I.; Chauzat, M.P.; Chen, Y.P.; Doublet, V.; Genersch, E.; Gisder, S.; Higes, M.; McMahon, D.P.; Martín-Hernández, R.; Natsopoulou, M.; et al. Standard methods for Nosema research. J. Apic. Res. 2013, 52, 1–28. [Google Scholar] [CrossRef] [Green Version]
  37. Bolger, A.M.; Lohse, M.; Usadel, B. Trimmomatic: A flexible trimmer for Illumina sequence data. Bioinformatics 2014, 30, 2114–2120. [Google Scholar] [CrossRef] [Green Version]
  38. Schmieder, R.; Edwards, R. Quality control and preprocessing of metagenomic datasets. Bioinformatics 2011, 27, 863–864. [Google Scholar] [CrossRef] [Green Version]
  39. Kolde, R.; Kolde, M.R. Package ‘pheatmap’. R Package 2015, 1, 790. Available online: https://CRAN.R-project.org/package=pheatmap (accessed on 3 August 2022).
  40. Oliveros, J. An interactive tool for comparing lists with Venn’s diagrams (2007–2015). 2018. Available online: https://bioinfogp.cnb.csic.es/tools/venny/index.html (accessed on 3 August 2022).
  41. Moriya, Y.; Itoh, M.; Okuda, S.; Yoshizawa, A.C.; Kanehisa, M. KAAS: An automatic genome annotation and pathway reconstruction server. Nucleic Acids Res. 2007, 35, W182–W185. [Google Scholar] [CrossRef] [Green Version]
  42. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCT method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef]
  43. R Core Team. R: A Language and Environment for Statistical Computing; R Version 4.0.3; R Foundation for Statistical Computing: Vienna, Austria, 2020. [Google Scholar]
  44. Wickham, H.; Navarro, D.; Pedersen, T. ggplot2: Elegant Graphics for Data Analysis; Springer: New York, NY, USA, 2016. [Google Scholar]
  45. Kassambara, A.; Kassambara, M.A. Package ‘ggpubr’. 2020. Available online: https://rpkgs.datanovia.com/ggpubr/ (accessed on 3 August 2022).
  46. Kassambara, A.; Kosinski, M.; Biecek, P.; Fabian, S. Package ‘Survminer’. Drawing Survival Curves Using ‘ggplot2’, R Package version 03.1. 2017. Available online: http://www.sthda.com/english/rpkgs/survminer/ (accessed on 3 August 2022).
  47. Moore, D.F. Applied Survival Analysis Using R.; Springer: Berlin/Heidelberg, Germany, 2016. [Google Scholar]
  48. Spss, I. IBM SPSS Statistics for Windows, version 20.0; IBM Corp: Armonk, NY, USA, 2011; Volume 440, p. 394. [Google Scholar]
  49. Fontbonne, R.; Garnery, L.; Vidau, C.; Aufauvre, J.; Texier, C.; Tchamitchian, S.; El Alaoui, H.; Brunet, J.L.; Delbac, F.; Biron, D.G. Comparative susceptibility of three Western honeybee taxa to the microsporidian parasite Nosema ceranae. Infect. Genet. Evol. 2013, 17, 188–194. [Google Scholar] [CrossRef]
  50. Paris, L.; El Alaoui, H.; Delbac, F.; Diogon, M. Effects of the gut parasite Nosema ceranae on honey bee physiology and behavior. Curr. Opin. Insect Sci. 2018, 26, 149–154. [Google Scholar] [CrossRef]
  51. Dussaubat, C.; Brunet, J.-L.; Higes, M.; Colbourne, J.K.; Lopez, J.; Choi, J.-H.; Martin-Hernandez, R.; Botias, C.; Cousin, M.; McDonnell, C.; et al. Gut pathology and responses to the microsporidium Nosema ceranae in the honey bee Apis mellifera. PLoS ONE 2012, 7, e37017. [Google Scholar] [CrossRef] [Green Version]
  52. Goblirsch, M.; Huang, Z.Y.; Spivak, M. Physiological and behavioral changes in honey bees (Apis mellifera) induced by Nosema ceranae infection. PLoS ONE 2013, 8, e58165. [Google Scholar] [CrossRef]
  53. Fan, Y.; Wang, J.; Yu, K.; Zhang, W.; Cai, Z.; Sun, M.; Hu, Y.; Zhao, X.; Xiong, C.; Niu, Q.; et al. Comparative Transcriptome Investigation of Nosema ceranae Infecting Eastern Honey Bee Workers. Insects 2022, 13, 241. [Google Scholar] [CrossRef]
  54. Heinz, E.; Hacker, C.; Dean, P.; Mifsud, J.; Goldberg, A.V.; Williams, T.A.; Nakjang, S.; Gregory, A.; Hirt, R.P.; Lucocq, J.M.; et al. Plasma membrane-located purine nucleotide transport proteins are key components for host exploitation by microsporidian intracellular parasites. PLoS Pathog. 2014, 10, e1004547. [Google Scholar] [CrossRef] [Green Version]
  55. Vidau, C.; Panek, J.; Texier, C.; Biron, D.G.; Belzunces, L.P.; Le Gall, M.; Broussard, C.; Delbac, F.; El Alaoui, H. Differential proteomic analysis of midguts from Nosema ceranae-infected honeybees reveals manipulation of key host functions. J. Invertebr. Pathol. 2014, 121, 89–96. [Google Scholar] [CrossRef]
  56. Vávra, J.; Ronny Larsson, J. Structure of the microsporidia. In The Microsporidia and Microsporidiosis; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 1999; pp. 7–84. [Google Scholar]
  57. He, Q.; Vossbrinck, C.R.; Yang, Q.; Meng, X.-Z.; Luo, J.; Pan, G.-Q.; Zhou, Z.-Y.; Li, T. Evolutionary and functional studies on microsporidian ATP-binding cassettes: Insights into the adaptation of microsporidia to obligated intracellular parasitism. Infect. Genet. Evol. 2019, 68, 136–144. [Google Scholar] [CrossRef]
  58. Nelson, N. Metal ion transporters and homeostasis. EMBO J. 1999, 18, 4361–4371. [Google Scholar] [CrossRef] [Green Version]
  59. Freisinger, E.; Sigel, R.K. From nucleotides to ribozymes—A comparison of their metal ion binding properties. Coord. Chem. Rev. 2007, 251, 1834–1851. [Google Scholar] [CrossRef] [Green Version]
  60. Sigel, R.; Sigel, H. Metal ion interactions with nucleic acids and their constituents. Compr. Inorg. Chem. II 2013, 3, 623–660. [Google Scholar]
  61. Schweins, T.; Wittinghofer, A. GTP-binding proteins: Structures, interactions and relationships. Curr. Biol. 1994, 4, 547–550. [Google Scholar] [CrossRef]
  62. Hochstrasser, M. Ubiquitin-dependent protein degradation. Annu. Rev. Genet. 1996, 30, 405–439. [Google Scholar] [CrossRef]
  63. Kim, I.H.; Kim, D.J.; Gwak, W.S.; Woo, S.D. Increased survival of the honey bee Apis mellifera infected with the microsporidian Nosema ceranae by effective gene silencing. Arch. Insect Biochem. Physiol. 2020, 105, e21734. [Google Scholar] [CrossRef]
  64. Yang, D.; Pan, L.; Chen, Z.; Du, H.; Luo, B.; Luo, J.; Pan, G. The roles of microsporidia spore wall proteins in the spore wall formation and polar tube anchorage to spore wall during development and infection processes. Exp. Parasitol. 2018, 187, 93–100. [Google Scholar] [CrossRef]
  65. Wu, Z.; Li, Y.; Pan, G.; Tan, X.; Hu, J.; Zhou, Z.; Xiang, Z. Proteomic analysis of spore wall proteins and identification of two spore wall proteins from Nosema bombycis (Microsporidia). Proteomics 2008, 8, 2447–2461. [Google Scholar] [CrossRef]
  66. Wu, Z.; Li, Y.; Pan, G.; Zhou, Z.; Xiang, Z. SWP25, a novel protein associated with the Nosema bombycis endospore 1. J. Eukaryot. Microbiol. 2009, 56, 113–118. [Google Scholar] [CrossRef]
  67. Ni, W.; Bao, J.; Mo, B.; Liu, L.; Li, T.; Pan, G.; Chen, J.; Zhou, Z. Hemocytin facilitates host immune responses against Nosema bombycis. Dev. Comp. Immunol. 2020, 103, 103495. [Google Scholar] [CrossRef]
  68. Waterhouse, A.; Bertoni, M.; Bienert, S.; Studer, G.; Tauriello, G.; Gumienny, R.; Heer, F.T.; de Beer, T.A.P.; Rempfer, C.; Bordoli, L. SWISS-MODEL: Homology modelling of protein structures and complexes. Nucleic Acids Res. 2018, 46, W296–W303. [Google Scholar] [CrossRef] [Green Version]
  69. Spagnolo, L.; Rivera-Calzada, A.; Pearl, L.H.; Llorca, O. Three-dimensional structure of the human DNA-PKcs/Ku70/Ku80 complex assembled on DNA and its implications for DNA DSB repair. Mol. Cell 2006, 22, 511–519. [Google Scholar] [CrossRef] [PubMed]
  70. Mazet, F.; Yu, J.-K.; Liberles, D.A.; Holland, L.Z.; Shimeld, S.M. Phylogenetic relationships of the Fox (Forkhead) gene family in the Bilateria. Gene 2003, 316, 79–89. [Google Scholar] [CrossRef]
  71. Partridge, L.; Brüning, J. Forkhead transcription factors and ageing. Oncogene 2008, 27, 2351–2363. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Ohi, M.D.; Link, A.J.; Ren, L.; Jennings, J.L.; McDonald, W.H.; Gould, K.L. Proteomics analysis reveals stable multiprotein complexes in both fission and budding yeasts containing Myb-related Cdc5p/Cef1p, novel pre-mRNA splicing factors, and snRNAs. Mol. Cell. Biol. 2002, 22, 2011–2024. [Google Scholar] [CrossRef] [Green Version]
  73. Kim, D.-U.; Maeng, S.; Lee, H.; Nam, M.; Lee, S.-J.; Hoe, K.-L. The effect of the cwf14 gene of fission yeast on cell wall integrity is associated with rho1. J. Microbiol. 2016, 54, 98–105. [Google Scholar] [CrossRef]
  74. Meier, U.T. The many facets of H/ACA ribonucleoproteins. Chromosoma 2005, 114, 1–14. [Google Scholar] [CrossRef] [Green Version]
  75. Mayes, A.E.; Verdone, L.; Legrain, P.; Beggs, J.D. Characterization of Sm-like proteins in yeast and their association with U6 snRNA. EMBO J. 1999, 18, 4321–4331. [Google Scholar] [CrossRef] [Green Version]
  76. Green, M.R. Pre-mRNA splicing. Annu. Rev. Genet. 1986, 20, 671–708. [Google Scholar] [CrossRef]
Figure 1. Observation and molecular identification of N. ceranae infection. (A) Kaplan–Meier survival curve (* = significant difference); (B) observation of mature spores; (C) mature spore counts (Different letters indicate significant differences (p < 0.05)); and (D) molecular detection of N. ceranae infection at 0.5 (12 h), 1, 3, 5, 10, 15, and 20 d.p.i.; d.p.i. = days post-infection; CD = cell debris; SP = spore; the arrowhead indicates the cell debris; the arrow indicates the spore.
Figure 1. Observation and molecular identification of N. ceranae infection. (A) Kaplan–Meier survival curve (* = significant difference); (B) observation of mature spores; (C) mature spore counts (Different letters indicate significant differences (p < 0.05)); and (D) molecular detection of N. ceranae infection at 0.5 (12 h), 1, 3, 5, 10, 15, and 20 d.p.i.; d.p.i. = days post-infection; CD = cell debris; SP = spore; the arrowhead indicates the cell debris; the arrow indicates the spore.
Insects 13 00716 g001
Figure 2. DEG and common gene identification. (A) Scatter plot of DEGs (red dots = upregulation and blue dots = downregulation); (B) Venn plot and fold-change number for the distribution of common DEGs.
Figure 2. DEG and common gene identification. (A) Scatter plot of DEGs (red dots = upregulation and blue dots = downregulation); (B) Venn plot and fold-change number for the distribution of common DEGs.
Insects 13 00716 g002
Figure 3. Dot plot of N. ceranae Gene Ontology (GO) term enrichment analysis, showing the enriched pathways with p < 0.05. (A) Upregulated DEGs; (B) downregulated DEGs. Each comparison is indicated at the bottom. The total number of regulated genes within the GO pathways selected on the dot plot are indicated in brackets. Colors indicate the p values from Fisher’s exact test, and the dot size is proportional to the number of DEGs and common DEGs in the pathway.
Figure 3. Dot plot of N. ceranae Gene Ontology (GO) term enrichment analysis, showing the enriched pathways with p < 0.05. (A) Upregulated DEGs; (B) downregulated DEGs. Each comparison is indicated at the bottom. The total number of regulated genes within the GO pathways selected on the dot plot are indicated in brackets. Colors indicate the p values from Fisher’s exact test, and the dot size is proportional to the number of DEGs and common DEGs in the pathway.
Insects 13 00716 g003
Figure 4. Dot plot of N. ceranae Gene Ontology (GO) term enrichment analysis, showing the enriched pathways with p < 0.05. (A) Core genes; (B) stage-specific genes. Each trial is indicated at the bottom; the total number of regulated genes within the GO pathways selected on the dot plot are indicated in brackets. Colors indicate the p values from Fisher’s exact test, and the dot size is proportional to the number of DEGs and common DEGs in the pathway.
Figure 4. Dot plot of N. ceranae Gene Ontology (GO) term enrichment analysis, showing the enriched pathways with p < 0.05. (A) Core genes; (B) stage-specific genes. Each trial is indicated at the bottom; the total number of regulated genes within the GO pathways selected on the dot plot are indicated in brackets. Colors indicate the p values from Fisher’s exact test, and the dot size is proportional to the number of DEGs and common DEGs in the pathway.
Insects 13 00716 g004
Figure 5. KEGG annotation of (AC) DEGs and (D) common DEGs. The bar plots show the number of DEGs identified in the relative pathway based on the KEGG pathway database.
Figure 5. KEGG annotation of (AC) DEGs and (D) common DEGs. The bar plots show the number of DEGs identified in the relative pathway based on the KEGG pathway database.
Insects 13 00716 g005
Figure 6. Validation of N. ceranae gene expression in upregulated common DEGs during N. ceranae infection. The bar plots show the relative gene expression levels of different common DEGs at different d.p.i. Different letters indicate significant differences (p < 0.05).
Figure 6. Validation of N. ceranae gene expression in upregulated common DEGs during N. ceranae infection. The bar plots show the relative gene expression levels of different common DEGs at different d.p.i. Different letters indicate significant differences (p < 0.05).
Insects 13 00716 g006
Figure 7. Validation of N. ceranae gene expression in upregulated stage-specific genes during N. ceranae infection. The bar plots show the relative gene expression levels of different stage-specific genes at different d.p.i. The black arrowheads indicate the time point at which the highest expression level of the gene was predicted based on RNA-seq data. Different letters indicate significant differences (p < 0.05).
Figure 7. Validation of N. ceranae gene expression in upregulated stage-specific genes during N. ceranae infection. The bar plots show the relative gene expression levels of different stage-specific genes at different d.p.i. The black arrowheads indicate the time point at which the highest expression level of the gene was predicted based on RNA-seq data. Different letters indicate significant differences (p < 0.05).
Insects 13 00716 g007
Table 1. List of upregulated common DEGs with FPKM values > 1000 at different times post-infection.
Table 1. List of upregulated common DEGs with FPKM values > 1000 at different times post-infection.
No.Gene IDFPKM
(Mature Spore/5 d.p.i/10 d.p.i/20 d.p.i.)
Fold ChangeGene TypeProtein IDLength (aa)Protein NameE-ValueKO
Annotation
NcSp/5 d.p.i.NcSp/10 d.p.i.NcSp/20 d.p.i.NCBIEMBL-EBI
1 *AAJ76_160005294315,626.3/115,903/34,433.6/47165.27.4171245742.2035658353.018307934Protein codingXP_024331363.1265Hypothetical proteinSpore wall protein 25 (SWP25_NOSB1)1.50 × 10−13-
2 *AAJ76_200014916112,535.3/91,333/34,160.1/43545.17.2860656792.7251204533.473788695Protein codingXP_024332331.1263Hypothetical protein---
3 *AAJ76_200014184511,759.8/65,562.7/29,751.5/34741.65.5751475992.5299169042.95426655Protein codingXP_024332326.1341Hypothetical proteinSpore wall protein 30 (SWP32_NOSB)6.00 × 10−13-
4AAJ76_1600051435189.533/21,504.1/1488.81/3201.29113.45843047.85515639716.89038139Protein codingXP_024331362.1286Hypothetical proteinSpore wall protein 25 (SWP25_NOSB)1.50 × 10−6-
5AAJ76_23000224186875.45/16,903.7/17,987.8/19,037.22.4585539272.6162311282.768875144Protein codingXP_024331106.1198Hypothetical proteinWD repeat-containing protein 48 homolog (WDR48_DICDI)3.30 × 10−5-
6AAJ76_8000115509783.185/14,628.1/1618.08/1750.2618.677765962.0660139562.234791131Protein codingXP_024331780.1331Hypothetical proteinRadixin (RADI_PIG)9.70 × 10−7-
7AAJ76_500093385836.136/12,300.3/3753.56/3436.6814.710965194.489173954.110203999Protein codingXP_024332003.1166Hypothetical protein---
8AAJ76_500085842830.632/8069.28/1893.56/1693.659.7146364032.2796621022.038997491Protein codingXP_024331997.1187Hypothetical protein---
9AAJ76_700014852292.943/6718.83/2035.24/2486.3722.935651766.9475773918.487560776Protein codingXP_024331793.1154Ubiquitin-conjugating enzyme e2 e2Ubiquitin-conjugating enzyme E2 E2 (UB2E2_MOUSE)1.10 × 10−60K20217
10 *AAJ76_500091146142.197/5518/1477.09/1142.3638.8055204810.387668698.033618765Protein codingXP_024332001.1198Forkhead hnf3 transcription factorFork head domain transcription factor slp1 (SLP1_DROME)6.30 × 10−25K09399
11AAJ76_1500044498794.305/5221.61/2185.54/1886.356.5738110962.751522262.374839476Protein codingXP_024331404.1233Proteasome subunit alpha type-5Probable proteasome subunit alpha type-5 (PSA5_ENCCU)8.10 × 10−98K02729
12AAJ76_1040006841612.16/4981.59/1332.9/1251.918.1377502252.1773751542.04506968Protein codingXP_024329929.1708cdc48-like aaa ATPase--K13525
13AAJ76_7900011594530.277/3735.63/2918.4/1222.327.0446596165.5035404162.305069101Protein codingXP_024330110.1199Peptidyl-prolyl cis-trans isomerasePeptidyl-prolyl cis-trans isomerase B (PPIB_RAT)1.00 × 10−40K03768
14AAJ76_5600010148487.495/2612.36/2107.93/1300.25.3587629564.3240104212.667114941Protein codingXP_024330347.1149Signal peptidase complex subunit 3Probable signal peptidase complex subunit 3 (SPCS3_CAEEL)3.30 × 10−5K12948
15AAJ76_2400023015423.143/2460.52/1559.48/1320.585.814882343.6854735393.120873758Protein codingXP_024331065.1229Proteasome subunit alpha type-2-aProteasome subunit alpha type-2 (PSA2_ORYSJ)3.40 × 10−36K02726
16AAJ76_3500034447417.662/1822.04/1577.38/2178.814.3624844843.7766837995.21667299Protein codingXP_024330780.120520s proteasome subunitProteasome subunit beta type-3 (PSB3_DICDI)1.50 × 10−47K02735
17AAJ76_1100054231704.144/1773.09/1670.11/1547.412.5180728112.3718289992.197586572Protein codingXP_024331572.1206Udp-glucose pyrophosphorylase---
* = selected gene for validation; FPKM = fragment per kilobase per million; d.p.i. = days post-infection; aa = amino acid; NCBI = National Center for Biotechnology Information; EMBL-EBI = the European Molecular Biology Laboratory European Bioinformatics Institute.
Table 2. List of downregulated common DEGs with FPKM values > 1000 at different times post-infection.
Table 2. List of downregulated common DEGs with FPKM values > 1000 at different times post-infection.
No.Gene IDFPKM
(Mature Spore/5 d.p.i/10 d.p.i/20 d.p.i.)
Fold Change Gene TypeProtein IDLength (aa)Protein NameE-ValueKO
Annotation
NcSp/5 d.p.i.NcSp/10 d.p.i.NcSp/20 d.p.i.NCBIEMBL-EBI
1AAJ76_630009567398,502/32947.1/61,754.1/70,676.1−12.0951772−6.453044615−5.6384163Protein codingXP_024330265.1107Hypothetical protein--K09140
2AAJ76_2000146453522,948/111,302/18,531.6/16,522.3−4.698434055−28.21909861−31.65101806Protein codingXP_024332329.1183Hypothetical proteinCilia- and flagella-associated protein 251 (CF251_MOUSE)2.00 × 10−7-
3AAJ76_1600023921978,848/95,231.2/24,913.3/19,641−10.27865792−39.29026518−49.83681619Protein codingXP_024331347.1267Hypothetical proteinNuclear protein 58 (NOP58_DICDI)2.40 × 10−20-
4AAJ76_1900025375321,105/84,821.6/14,069.3/17,252.2−3.785647288−22.8230548−18.61237171Protein codingXP_024331241.1275Polar tube protein 2Polar tube protein 2 (PTP2_ENCCU)2.60 × 10−6-
5AAJ76_1900028347282,631/79,179.8/8975.75/7515.33−3.569495047−31.4884317−37.60727843Protein codingXP_024331242.1446Hypothetical proteinProline-rich protein 2 (PRP2_MOUSE)3.10 × 10−22K10844
6AAJ76_1200007387104,357/36,502.7/9664.27/6697.23−2.858899688−10.79820795−15.58212405Protein codingXP_024329835.1464Elongation factor 1Elongation factor 1-alpha (EF1A_ENCCU)1.60 × 10−168K03857
7AAJ76_6000015274135,745/34,449.1/7043/6780.66−3.940448921−19.27367797−20.01944398Protein codingXP_024330295.1236Hypothetical protein---
8AAJ76_450001793180,571.7/30,587.7/9618.05/9332.46−2.634118045−8.377152866−8.633466829Protein codingXP_024330533.1239Hypothetical protein---
9AAJ76_21000240159,830.8/26,523.1/6006.5/3556.09−2.255801308−9.960995968−16.82494636Protein codingXP_024331167.1678Heat shock protein 70---
10AAJ76_130006800105,634/24,439/2415.33/1928.25−4.322332328−43.73480633−54.78199877Protein codingXP_024331464.1204Hypothetical protein---
11AAJ76_310001585854,826.6/22,173.9/2287.82/2631.64−2.47256766−23.96455092−20.83359639Protein codingXP_024330869.1338Hypothetical proteinUncharacterized protein ECU05_0110 (Y511_ENCCU)1.60 × 10−7-
12AAJ76_42000490891,596.5/21,857.3/5018.44/4703.25−4.190667331−18.25196643−19.47511578Protein codingXP_024330583.1221Hypothetical proteinSpore wall protein ECU02_0150 (Y215_ENCCU)6.70 × 10−25K01689
13AAJ76_270004308539,025.5/19,267.6/3836.51/3059.65−2.02544653−10.17213071−12.7549028Protein codingXP_024330997.1156Hypothetical protein---
14AAJ76_120004487684,245.8/18,063.1/3850.84/6236.68−4.663974821−21.87725765−13.50807323Protein codingXP_024331532.1222Spore wall protein precursorSpore wall protein 26 (SWP26_NOSB1)5.50 × 10−29-
15AAJ76_470002460367,259.6/16,069.1/3500.31/3049.83−4.185645126−19.2153853−22.05356525Protein codingXP_024330499.1284ABC-type multidrug transport ATPase and permease componentSpore wall protein 7 (SWP7_NOSB1)2.80 × 10−100-
16AAJ76_660001359834,656.4/16,003.4/2672.38/2821.32−2.165559782−12.96832278−12.28375723Protein codingXP_024330229.1317Hypothetical proteinTranscription initiation factor TFIID subunit 3 (TAF3_CHICK)3.10 × 10−6-
17AAJ76_140003676195,516.2/14,664.3/5710.68/3838.4−6.513531706−16.72587754−24.88440523Protein codingXP_024331438.1226Hypothetical protein---
18AAJ76_240003216938,208.9/13,393.2/2471.53/3059.12−2.852862075−15.45958836−12.49015954Protein codingXP_024331074.1131Hypothetical proteinHistone-lysine N-methyltransferase, H3 lysine-79 specific (9DOT1L_DICDI)1.10 × 10−6K07178
19AAJ76_530002263039,320.1/12,093.2/3489.63/3466.58−3.251442701−11.267691−11.34260414Protein codingXP_024330400.1295Hydrolase of the alpha beta-hydrolaseUncharacterized protein ECU07_0920 (Y792_ENCCU)1.60 × 10−37-
20AAJ76_100006589996,438/11,363/4552.52/4869.86−8.487031311−21.18351365−19.80303982Protein codingXP_024331635.1216Hypothetical protein---
21AAJ76_4000112883117,599/9836.2/2044.46/1175.49−11.95572738−57.52072681−100.0425551Protein codingXP_024332099.1327glyceraldehyde-3-phosphate dehydrogenaseGlyceraldehyde-3-phosphate dehydrogenase (G3P_CUTCU)2.10 × 10−120K00134
22AAJ76_500010974921,356.7/9085.52/2351.82/2527.34−2.350634122−9.080929654−8.450285003Protein codingXP_024332015.1229Spore wall protein 12Spore wall protein 12 (SWP12_NOSB1)2.20 × 10−76-
23AAJ76_230003639821,838.9/8533.66/3482.57/2535.92−2.559141644−6.270932541−8.611771259Protein codingXP_024331119.1380Spore wall protein 9Spore wall protein 9 (SWP9_NOSB1)2.00 × 10−39K07955
24AAJ76_270003177926,552.2/7739.67/1271.69/1831.42−3.430669148−20.87942389−14.498181Protein codingXP_024330986.1167Hypothetical protein---
25AAJ76_660001972828,729.1/5107.81/2222.85/3656.61−5.624520074−12.9244412−7.856735542Protein codingXP_024330233.1154Prenylated Rab acceptor 1-like proteinUncharacterized membrane protein ECU02_1470 (Y2E7_ENCCU)3.70 × 10−16K11838
26AAJ76_42000413418,934/5107.54/1904.29/1739.95−3.707071651−9.942853842−10.88191078Protein codingXP_024330582.1219Ran-specific GTPase-activating protein 1Ran-specific GTPase-activating protein 1 (YRB1_ENCCU)3.70 × 10−41-
27AAJ76_20005984819,359.4/4708.45/1848.74/1879.72−4.111628734−10.47167266−10.29905455Protein coding XP_024332275.1250Hypothetical proteinGelsolin-related protein of 125 kDa (GNRA_DICDI)3.90 × 10−5-
28AAJ76_160006850214,848.7/4309.26/2652.73/2834.43−3.445778582−5.597528622−5.238704187Protein codingXP_024331374.1122ccaat binding transcription factor subunit aNuclear transcription factor Y subunit B-5 (NFYB5_ARATH)1.20 × 10−24-
29AAJ76_1100066671659,735/4001.88/4268.18/8495.99−164.8568473−154.570372−77.65247484Protein codingXP_024331582.1132Hypothetical proteinProtein DR_1172 (UB72_DEIRA)8.90 × 10−15K11204
30AAJ76_100014258518,723.8/3995.19/1946.34/2392.26−4.686594587−9.620020711−7.826840255Protein codingXP_024332442.1268SeptinCell division control protein 11 (CDC11_ENCCU)4.30 × 10−98K01151
31AAJ76_28000673332,568.2/3946.19/1832.71/1589.86−8.253064983−17.77059399−20.48491498Protein codingXP_024330938.1 201Hypothetical protein---
32AAJ76_170006250319,005.2/2947.71/7507.76/2855.59−6.447411209−2.531407907−6.655421327non coding RNA--23S ribosomal RNA ---
33AAJ76_71000857222,045.4/2941.14/1179.25/1501.71−7.495527438−18.69447434−14.68010111Protein codingXP_024330184.1190Hypothetical protein--K00162
34AAJ76_10000808173,487.3/2295.2/1218.66/2315.53−32.01774949−60.30141705−31.73669519Protein codingXP_024331599.1209Hypothetical protein---
35AAJ76_10001650538466.55/1740.59/1290.67/1686.56−4.864191406−6.559791681−5.020009172Protein codingXP_024332456.1347Septin-like proteinCell division control protein 10 (CDC10_ENCCU)8.30 × 10−110-
36AAJ76_100001620597,815.6/1570.42/2985.17/2356.31−62.28660354−32.76718881−41.51209968Protein codingXP_024331606.1186Hypothetical protein---
FPKM = fragment per kilobase per million; d.p.i. = days post-infection; aa = amino acid; NCBI = National Center for Biotechnology Information; EMBL-EBI = the European Molecular Biology Laboratory European Bioinformatics Institute.
Table 3. List of the top 10 ranked stage-specific genes at different times post-infection.
Table 3. List of the top 10 ranked stage-specific genes at different times post-infection.
StageNo.Gene IDFPKMProtein IDLength (aa)Protein NameE-ValueKO Annotation
NCBIEMBL-EBI
5 d.p.i1 *AAJ76_50000264852388.86XP_024330451.1116g10 proteinPre-mRNA-splicing factor cwf14 (CWF14_SCHPO)7.40 × 10−27-
2AAJ76_5000723121067.07XP_024331989.1149Hypothetical protein---
3AAJ76_4500025263987.901XP_024330536.1176Hypothetical protein---
4AAJ76_500090033754.257XP_024332000.1164Nuclear essential protein 1Ribosomal RNA small subunit methyltransferase mra1 (NEP1_SCHPO)8.40 × 10−11K14568
5AAJ76_4800030660679.592XP_024330487.1132Hypothetical protein---
6AAJ76_2700033485549.13XP_024330988.1174Hypothetical protein---
7AAJ76_2200018137404.002XP_024331140.1145Deoxycytidylate deaminasetRNA-specific adenosine deaminase 2 (ADAT2_XENTR)8.60 × 10−22K15441
8AAJ76_900062461401.31XP_024331679.1148Bis (5 -adenosyl)-triphosphatase-like proteinBis(5’-adenosyl)-triphosphatase (FHIT_DICDI)1.00 × 10−25K01522
9AAJ76_1000019447326.831XP_024331608.1187Hypothetical protein---
10AAJ76_1400060351208.221XP_024331456.1199Hypothetical proteinPutative chaperone PpiD (PPID_BUCBP)1.60 × 10−5-
10 d.p.i1AAJ76_28500022565462.22XP_024329453.141Hypothetical protein---
2AAJ76_29000153003735.49XP_024330918.141Hypothetical protein---
3AAJ76_11000102653263.98XP_024331553.143Hypothetical protein---
4 *AAJ76_24000201902691.71XP_024331062.153H aca ribonucleoprotein complex subunit 3Putative H/ACA ribonucleoprotein complex subunit 3 (NOP10_CAEEL)8.80 × 10−11-
5AAJ76_9000279411915.56XP_024331657.163Hypothetical protein---
6AAJ76_11000541911914.4XP_024331571.156Hypothetical proteinExportin-T (XPOT_DICDI)4.80 × 10−4-
7AAJ76_21600031472.66XP_024329567.162Nol1 nol2 sun tRNA rRNA cytosine-c5-methylasetRNA (cytosine(48)-C(5))-methyltransferase (TRM4_METJA)2.80 × 10−12K14835
8AAJ76_11200010851441.18XP_024329876.152Hypothetical protein---
9AAJ76_36000237781344.19XP_024330748.163Hypothetical protein---
10AAJ76_10001663131281.3XP_024332457.164Hypothetical protein---
20 d.p.i1AAJ76_30001325345797.14XP_024332206.146Hypothetical protein---
2 *AAJ76_23000354675519.29XP_024331118.167Hypothetical proteinU6 snRNA-associated Sm-like protein LSm6 (LSM6_SCHPO)8.70 × 10−7-
3AAJ76_61000160642662.52XP_024330287.153Hypothetical protein---
4AAJ76_6000740611545.98XP_024331899.187Zinc finger domain-containing proteinPre-mRNA-splicing factor CWC24 (CWC24_KLULA)2.70 × 10−10K13127
5AAJ76_3470001881394.2XP_024329388.162Hypothetical protein---
6AAJ76_17200051461258.07XP_024329669.161Calcium-binding protein of the recoverin subfamily Neuronal calcium sensor 1 (NCS1_LYMST)3.60 × 10−9K19932
7AAJ76_4000289411057.47XP_024332054.186Ubiquitin-related modifier 1Ubiquitin-related modifier 1 (URM1_LACBS)5.40 × 10−7-
8AAJ76_230009864789.777XP_024331097.182Hypothetical protein---
9AAJ76_590008095769.682XP_024330308.184Hypothetical protein---
10AAJ76_610001724.872XP_024330279.169Hypothetical protein---
d.p.i.= days post-infection; * = selected gene for validation; FPKM = fragment per kilobase per million; aa = amino acid; NCBI = National Center for Biotechnology Information; EMBL-EBI = the European Molecular Biology Laboratory European Bioinformatics Institute.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Li, Y.-H.; Chang, Z.-T.; Yen, M.-R.; Huang, Y.-F.; Chen, T.-H.; Chang, J.-C.; Wu, M.-C.; Yang, Y.-L.; Chen, Y.-W.; Nai, Y.-S. Transcriptome of Nosema ceranae and Upregulated Microsporidia Genes during Its Infection of Western Honey Bee (Apis mellifera). Insects 2022, 13, 716. https://doi.org/10.3390/insects13080716

AMA Style

Li Y-H, Chang Z-T, Yen M-R, Huang Y-F, Chen T-H, Chang J-C, Wu M-C, Yang Y-L, Chen Y-W, Nai Y-S. Transcriptome of Nosema ceranae and Upregulated Microsporidia Genes during Its Infection of Western Honey Bee (Apis mellifera). Insects. 2022; 13(8):716. https://doi.org/10.3390/insects13080716

Chicago/Turabian Style

Li, Yi-Hsuan, Zih-Ting Chang, Ming-Ren Yen, Yu-Feng Huang, Tzu-Han Chen, Ju-Chun Chang, Ming-Cheng Wu, Yu-Liang Yang, Yue-Wen Chen, and Yu-Shin Nai. 2022. "Transcriptome of Nosema ceranae and Upregulated Microsporidia Genes during Its Infection of Western Honey Bee (Apis mellifera)" Insects 13, no. 8: 716. https://doi.org/10.3390/insects13080716

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop