Skip to main content
  • Research article
  • Open access
  • Published:

Early transcriptome changes induced by the Geminivirus C4 oncoprotein: setting the stage for oncogenesis

Abstract

Background

The Beet curly top virus C4 oncoprotein is a pathogenic determinant capable of inducing extensive developmental abnormalities. No studies to date have investigated how the transcriptional profiles differ between plants expressing or not expressing the C4 oncoprotein.

Results

We investigated early transcriptional changes in Arabidopsis associated with expression of the Beet curly top virus C4 protein that represent initial events in pathogenesis via a comparative transcriptional analysis of mRNAs and small RNAs. We identified 48 and 94 differentially expressed genes at 6- and 12-h post-induction versus control plants. These early time points were selected to focus on direct regulatory effects of C4 expression. Since previous evidence suggested that the C4 protein regulated the brassinosteroid (BR)-signaling pathway, differentially expressed genes could be divided into two groups: those responsive to alterations in the BR-signaling pathway and those uniquely responsive to C4. Early transcriptional changes that disrupted hormone homeostasis, 18 and 19 differentially expressed genes at both 6- and 12-hpi, respectively, were responsive to C4-induced regulation of the BR-signaling pathway. Other C4-induced differentially expressed genes appeared independent of the BR-signaling pathway at 12-hpi, including changes that could alter cell development (4 genes), cell wall homeostasis (5 genes), redox homeostasis (11 genes) and lipid transport (4 genes). Minimal effects were observed on expression of small RNAs.

Conclusion

This work identifies initial events in genetic regulation induced by a geminivirus C4 oncoprotein. We provide evidence suggesting the C4 protein regulates multiple regulatory pathways and provides valuable insights into the role of the C4 protein in regulating initial events in pathogenesis.

Background

Plant virus proteins are adept at co-opting cellular machinery and metabolic pathways to alter the host physiology to benefit the virus life cycle [1]. One such virus protein is the small C4 protein (~ 10 kDa) (AC4 in geminiviruses with bipartite genomes) encoded by some members of the Geminiviridae [2]. Viruses within the Geminiviridae family cause a variety of economically important diseases in crop plants worldwide [3]. Studies on the role of the C4/AC4 proteins from members of the Curtovirus and Begomovirus genera suggest that the proteins have a diverse set of functions by which they modulate pathogenesis.

Some geminivirus C4/AC4 proteins have been shown to induce oncogenesis during virus infection [4] and when expressed ectopically [4,5,6,7,8]. The oncogenic nature of C4/AC4 proteins has been shown to, at least in part, result from their ability to interfere with the function of shaggy-like protein kinases [8,9,10]. In Arabidopsis, seven shaggy-like protein kinases (AtSKs) of the ten-member multigene family have been implicated in negatively regulating the brassinosteroid (BR) signaling pathway (BRSP) [11,12,13,14]. BR is a steroid hormone that functions as a master regulator of plant development, growth and adaption to stress [15]. AtSKs have also been implicated in crosstalk between the BRSP and other hormone signaling pathways, biotic and abiotic stress responses, root and stomata development, flower development, xylem differentiation, phloem development, and pattern-triggered immunity [13, 14, 16,17,18,19].

AtSKs regulate the closely related BRI1-EMS SUPPRESSOR 1 (BES1) and BRASSINAZOLE-RESISTANT 1 (BZR1) transcription factors, which are pivotal in the BRSP [14]. In the absence of BR, AtSKs hyperphosphorylate and inactivate BES1 and BZR1. In the presence of BR, the hormone binds to the cell surface receptor kinase BRASSINOSTEROID INSENSITIVE 1 (BRI1) and co-receptor BRASSINOSTEROID INSENSITIVE 1-ASSOCIATED RECEPTOR KINASE 1 (BAK1), initiating a signaling cascade that results in the negative regulation of AtSKs and the activation of BES1/BZR1 transcription factors. BZR1/BES1 regulate a large number of BR-responsive genes and repress BR biosynthetic genes [15, 20,21,22,23,24,25]. Activated BES1/BZR1 transcription factors coordinate a complex multisignal regulatory network controlling growth and development [26].

The C4 protein of the curtovirus Beet curly top virus (BCTV) binds to the 7 AtSKs implicated in BR signaling. The protein interferes with the function of the AtSKs and sequesters the kinases to the plasma membrane (PM; 9). Chemical inhibition of the same 7 AtSKs with bikinin induces hyperplasia that phenocopies symptoms induced by C4, suggesting hyperplasia may, at least in part, be due to C4 modulating the function of some combination of the 7 AtSKs [9]. Consistent with this, the C4 protein of the begomovirus Tomato leaf curl Yunnan virus (TLCYnV) also was shown to interact with N. benthamiana shaggy-like protein kinase η (NbSKη) and sequester the kinase to the PM [8]. This interaction was suggested to impair NbSKη directed degradation of NbCycD1;1, resulting in abnormal cell division [8]. Interactions between three additional begomovirus C4 proteins and shaggy-like protein kinases have also been confirmed although their role in pathogenesis is not known [10, 27, 28].

A number of other C4/AC4-host protein interactions have been identified. A curtovirus C4 protein was shown to bind to CLAVATA 1 (CLV1) [29]. Two begomovirus C4/AC4 proteins were shown to interact with CLV1-type PM receptor-like kinases BARELY ANY MERISTEM 1 and 2 (BAM1, BAM2) [30, 31], interfering with their ability to regulate cell-to-cell movement of RNAi [30]. CLV1, BAM1 and BAM2 are required for shoot apical meristem homeostasis, as well as vascular tissue, anther and root development [32]. In addition, begomovirus C4 proteins have been shown to interact with S-ADENOSYL METHIONINE SYNTHETASE and AGONAUTE 4, proteins that modulate gene silencing [33, 34], and HYPERSENSITIVE INDUCED REACTION 1 (HIR1), impairing the HIR1-mediated hypersensitive response [35].

Transcriptional analyses of geminivirus-infected plants have been shown to impact defense/immune responses, hormone homeostasis, the cell cycle and autophagy [36,37,38,39]. Ectopically expressed BCTV C4 leads to a severe developmental phenotype characterized by the loss of meristem function, prolific cell division and loss of cell-type differentiation [6]. These ectopically expressed changes are similar to those observed in the vascular tissue of infected plants, suggesting that ectopically expressed C4 recapitulates C4 pathogenesis observed during infection. To develop a better understanding of the role of the C4 protein in pathogenesis in the absence of BCTV infection, we performed a comparative transcriptional analysis from transgenic Arabidopsis plants expressing the BCTV C4 protein under the regulatory control of an inducible promoter relative to non-induced plants. To identify changes in C4-induced gene expression that are expected to represent initial events in pathogenesis, we chose early times post-C4 induction (6 and 12 h) as opposed to later times where more complex gene expression patterns likely would be composed of both direct and indirect changes to the transcriptome.

We observed that C4-induced transcriptional changes disrupted hormone homeostasis that were indicative of the protein regulating the BRSP. Other transcriptome changes suggest that C4 interferes with cell development, cell wall homeostasis, redox homeostasis and lipid transport in a C4-induced BRSP-independent manner. The results provide insights into the multifunction role of the BCTV C4 protein in virus infection and pathogenesis.

Results

Transcriptional analysis design

To begin understanding gene expression changes induced by the BCTV C4 protein, we performed RNA-seq analysis on RNA extracted from seedlings of Arabidopsis line IPC4–28 expressing or not expressing the C4 protein at 6- and 12-h post-induction (hpi) or post-mock induction (Fig. 1a). IPC4–28 is a transgenic line that expresses the BCTV C4 protein under the regulatory control of a ß-estradiol (ß-est) inducible promoter. Induction with ß-est results in near synchronized expression of the C4 protein, which is detectable by Western blot analysis as early as 6-hpi [6]. Initial symptoms of an abnormal development phenotype occurred in IPC4–28 seedlings geminated in liquid media in the presence of 10 μM ß-est as early as 2-days post-induction [6]. The noninduced IPC4–28 seedlings were unaltered and phenotypically identical to the wild type (WT) seedlings under light microscopy looking at cotyledon development and light microscopy and scanning electron microscopy looking at shoot and root apical meristem development [6]. We hypothesized that early effects on host gene expression caused by C4-host protein interactions would give insights into initial events leading to C4 induced pathogenesis. This would minimize complexities resulting from secondary down-stream gene expression changes that result with time following C4 expression or that develop from expression of other viral proteins during virus infection. RNA-seq resulted in a total of 576.0 million reads that passed quality control for 30 libraries, with an average of 19.2 million reads per library (Supplementary Fig. S1, Additional file 1). Of these reads, 85–88% (three library replicates/treatment) mapped to the Arabidopsis reference genome.

Fig. 1
figure 1

a Schematic of the experimental design and comparison of the number of differentially expressed up- (up arrows) and down-regulated (down arrows) genes at 6- and 12-h post-induction (hpi) of the BCTV C4 gene. Transgenic line IPC4–28 expresses the Beet curly top virus C4 gene under regulatory control of a ß-estradiol (ß-est) inducible promoter. Wild Type, Sei-O. Ind, Induced. Nonind, Noninduced. b Venn diagram showing the number of differentially expressed genes under different experimental conditions. C4_I_6_0 (blue), number of C4-induced differentially expressed genes at 6-hpi versus 0-hpi. C4_I_12_0 (orange), number of C4-induced differentially expressed genes at 12-hpi versus 0-hpi. Overlap (purple), number of C4-induced differentially expressed genes at 6-and 12-hpi versus 0-hpi

Differentially expressed genes following C4 induction

At 6-hpi, 48 DE genes were detected in induced IPC4–28 seedlings with 43 up-regulated (90%) and 5 down-regulated (10%). At 12-hpi, 94 DE genes were detected in induced IPC4–28 seedlings with 71 up-regulated (76%) and 23 down-regulated (24%) (Fig. 1b, Additional file 2). Six DE genes were distinct in seedlings at 6-hpi, 52 were distinct at 12-hpi and 42 were common to both time points (Fig. 1b). Of the DE genes in common at 6-hpi and 12-hpi, 37 (88%) were up regulated and expressed at 6-hpi > 78% of the levels detected at 12-hpi. Five were down regulated and repressed at 6-hpi > 84% of the levels detected at 12-hpi. Therefore, regulation of transcriptional differences detected at 6-hpi were generally maintained and amplified at 12-hpi. In control experiments with Sei-0 wild-type seedlings, no ß-est induced DE genes were detected in WT seedlings when compared to noninduced WT seedlings at 6- and 12-hpi (Fig. 1a), indicating that ß-est did not induce any detectable DE genes under the parameters used. Similarly, the only DE genes detected when comparing induce IPC4–28 seedlings to induced WT seedlings in the presence of 10 μM ß-est were the same DE genes detected when comparing induced IPC4–28 seedlings to noninduced IPC4–28 seedlings. To better understand the transcriptional dynamics from 0- through 12-hpi, a heatmap showing changes in the expression patterns of the DE genes is shown in Fig. 2. DE genes fell into 2 major clusters based on their patterns of expression and were distinct for induced IPC4–28 seedlings at 6- and 12-hpi relative to mock-induced IPC4–28 seedlings or induced and mock-induced wild-type seedlings. The group represented by induced IPC4–28 seedlings was composed of 2 subgroups separately representing induced IPC4–28 seedings at 6-hpi and at 12-hpi. This finding was supported by the principle component analysis (PCA) illustrated in Supplementary Fig. S2 (Additional file 1), which shows that replicates from 6- and 12-hpi transgenic induced samples are clustered at a distance from all other replicates, indicating a large variance in the expression of these two conditions comparing to the others.

Fig. 2
figure 2

Gene heat map showing hierarchical clustering of differentially expressed genes based on color-coded expression levels. Conditions indicate seedlings that were induced or noninduced at 6- and 12-hpi. C4 Trans, IPC4–28 seedlings. WT, Sei-0 seedlings. I, induced. NI, noninduced. Cluster in black oval represent IPC4–28 seedlings induced at 6- and 12-hpi. Subgroup in red  oval represents IPC4–28 seedlings induced at 6-hpi. Subgroup in blue  oval represents IPC4–28 seedlings induced at 12-hpi. The expression values are represented in the Log2 scale of normalized counts of gene expression

The number of C4 mRNA reads in non-induced IPC4–28 seedlings at 0-, 6-, and 12-hpi were at trace levels compared to levels in induced IPC4–28 seedlings (Additional file 2). For example, there was an average of 11,349 C4 mRNA reads in induced IPC4–28 samples at 12-hpi and an average of 22 C4 mRNA reads in noninduced IPC4–28 samples at 12 hpmi. No DE genes were detected under our parameters that would have resulted from trace amounts of C4 mRNA due to promoter leakage when comparing noninduced IPC4–28 and noninduced WT samples. However, we cannot absolutely exclude the possibility that some trace amount of C4 is present in noninduced IPC4–28 plants that might have a small effect on the transcriptome and was nondetectable under the parameters used.

Validation of differentially expressed genes detected by RNA-seq

To verify the RNA-seq data, 10 DE genes (4 down and 6 up-regulated) were validated by reverse transcription-quantitative PCR (RT-qPCR). The 4 down-regulated genes were: AT5G04950 (NICOTIANAMINE SYNTHASE 1, NAS1), AT1G05250 (PEROXIDASE 2, PRX2), AT5G46890 (lipid transfer protein) and AT2G47540 (POLLEN OLE E 1 ALLERGEN extensin family protein). The 6 up-regulated genes were: AT3G50770 (CALMODULIN-LIKE 41 PROTEIN, CML41), AT3G57240 (Β-1,3-GLUCANASE 3), AT5G65800 (1-AMINOCYCLOPROPANE-1-CARBOXYLATE SYNTHASE 5, ACS5), AT2G30770 (CYTOCHROME P71A13), AT5G25190 (ETHYLENE-RESPONSE TRANSCRIPTION FACTOR 003), AT2G44130 (KELCH-DOMAIN CONTAINING F-BOX PROTEIN 39). AT5G62690 (TUBULIN βCHAIN 2, TUB2) was also analyzed as a control gene not regulated by ß-est. Fold changes were compared for the 11 genes from RNA-Seq and RT-qPCR (Fig. 3). A Pearson correlation coefficient of R2 = 0.95, P < 0.0001) indicates a high correlation between the two methods and validates the RNA-seq data.

Fig. 3
figure 3

Validation of RNA-seq expression results by comparison with quantitative reverse transcription-PCR (RT-qPCR) data from the same RNA samples. Comparison of Log2 fold changes of 10 genes at 12-hpi. TUB2 was used as a control that was not differentially expressed in the RNA-seq experiments. MON1 was used as an endogenous housekeeping control. C4-induced transcriptional changes were normalized to gene expression in noninduced seedlings. Columns represent Log2 fold changes from RT-qPCR between ß-est treated and mock treated samples. Numbers in parenthesis indicate induced Log2 fold changes from RNA-seq expression data (Additional file 2) for direct comparison with RT-qPCR data. Log2 fold changes of 4 of the 10 genes that were differentially expressed at 6-hpi are also shown. Error bars indicate standard deviations from three biological replicates

C4 regulated DE genes that were BR-signaling pathway dependent and BR-signaling pathway independent

To determine the extent to which C4 regulates genes in the BRSP by inhibiting AtSKs, we compared the C4-responsive DE genes identified at 12-hpi with databases of BR-responsive genes and DE genes in bes1-D and bzr1-1D gain-of-function mutants (24, 25, Table 1 and Additional file 3). Bes1-D and bzr1-1D are constitutively active mutants [21, 22]. At 12-hpi, 43 (61%) of the C4 up-regulated DE genes were previously shown to be responsive to BR and/or one or both of the gain-of-function mutants, while 28 (39%) of the up-regulated DE genes were responsive to C4, but not to BR, bes1-D and/or bzr1-1D. Of the 23 down-regulated DE genes responsive to C4 at 12-hpi, 20 (87%) were regulated independent of the BRSP. Of these 20 C4 down-regulated DE genes, 4 have previously been shown to be up-regulated by BR and/or the bzr1-1D gain-of-function mutant and 16 were only responsive to C4. The results indicate that expression of C4 has a direct effect on regulating the BRSP (BRSP-dependent response). However, a subset of C4-regulated DE genes are uniquely responsive to C4 and independent of the BRSP (BRSP-independent).

Table 1 Number of C4-regulated genes induced in a brassinosteroid-signaling pathway dependent or independent manner at 12-hpi

Gene ontology analysis

To gain insights into biological processes the C4 protein modulates at early time points following induction, gene ontology (GO) enrichment analysis was utilized to identify GO terms over- or under-represented for DE genes identified in induced IPC4–28 seedlings at 6- and 12-hpi (Fig. 4). At 6-hpi, GO enrichment identified 13 categories from up-regulated DE genes involved in biological processes, including 8 associated with hormone-related processes. GO enrichment identified 2 related categories from down-regulated DE genes associated with lipid transport and lipid binding. Of the DE genes identified at 6-hpi, 65% were captured by GO enrichment categories. At 12-hpi, GO enrichment identified 17 categories from up-regulated DE genes, including 10 associated with hormone-related processes and 3 associated with pathogen defense. Five GO categories were identified from down-regulated DE genes, including 3 associated with cell development and 1 each associated with lipid transfer and cell wall processes (Fig. 4). At 12-hpi, 52% of the DE genes are members in GO enrichment categories.

Fig. 4
figure 4

GO categories representing over or under-represented genes differentially expressed in C4 expressing seedlings relative to non-expressing seedlings. Bars represent the percentage of regulated genes in addition to those expected by chance in the Arabidopsis reference (Enrichment test). GO categories were organized into those representing up-regulated genes or those representing down-regulated genes. Blue bars indicate gene numbers and orange bars indicated fold enrichment. Categories enclosed in green rectangles are hormone related, those enclosed in the red rectangle are defense/immune related and those enclosed in the grey rectangle are cell development related. All categories had false discovery rates with P < 0.05

Early events in C4 expression associated with hormone homeostasis

C4 up-regulated DE genes associated with hormone homeostasis, included genes involved in hormone biosynthesis and the regulation of hormone levels as well as DE genes responsive to auxin, BR and salicylic acid (Additional file 4 and Table 2). At 6- and 12-hpi, 40% (19 of 48) and 24% (23 of 94), respectively, of all DE genes were associated with hormone homeostasis. This indicates that much of the initial response to C4 resulted in up-regulated expression of genes that have an effect on hormone homeostasis or respond to alterations in hormone homeostasis, a trend seen at 6-hpi and increased at 12-hpi. All DE genes affecting hormone homeostasis at 12-hpi were responsive to BR and/or one or both of the bes1-D and bzr1-1D gain-of-function mutants with the exception of 1-AMINOCYCLOPROPANE-1-CARBOXYLATE SYNTHASE 9 (ACS9), LONELY GUY 5 (LOG5) and PARAXANTHINE METHYLTRANSFERASE 1 (PXMT1) (Table 2 and Additional file 3). The expression of the ABC TRANSPORTER G FAMILY 40 protein (ABCG40) was differentially responsive to C4 in that the gene is down-regulated in bes1-D but up-regulated following C4 expression. These results suggest that most of the hormone-related responses can be contributed to the role of the C4 protein in regulating the BRSP.

Table 2 Gene ontology (GO) terms indicating over-represented, up-regulated genes in hormone homeostasis at 12-h post-induction

Early events in C4 expression associated with host defense and redox homeostasis

The up-regulation of DE genes under the GO term response to salicylic acid observed at 12-hpi may suggest initial events in a salicylic acid-directed pathogen-related response induced by C4 (Table 2). In addition, GO analysis identified up-regulated DE genes associated with host defense (Table 3). Five of the 10 DE genes were significantly up-regulated at 6 hpi. Four additional DE genes that respond to pathogen infection were upregulated, but not captured by the GO enrichment analysis; CML41 (AT3G50770, facilitates rapid callose deposition at plasmodesmata when upregulated by bacterial infection [40];, FAD-LINKED OXIDOREDUCTASE 1 (AT1G26380, involved in biosynthesis of a cyanogenic phytoalexin and distinct from the canonical camalexin pathway, required for inducible pathogen defense [41];, METALLOENDOPROTEINASE 3-MMP (AT1G24140, contributes to pattern-triggered immunity by fungi in Arabidopsis [42]; and POLYGALACTURONASE-INHIBITING PROTEIN 2 (AT5G06870, an inhibitor of fungal polygalacturonase [43];. Nine of the 14 DE genes are up-regulated by BR and/or one or both bes1-D and bzr1-1D gain-of-function mutants (Additional file 3). Three DE genes were uniquely responsive to C4, INDOLEACETALDOXIME DEHYDRATASE (AT2G30770, cytochrome P450), AT2G43580 (an endochitinase), and PXMT 1. Interestingly, 2 DE genes, CML41 and ABCG40, were differentially responsive to C4 relative to the BRSP, being up-regulated following C4 expression, but down-regulated in bes1-D (Additional file 3). The only pathogenesis-related protein responsive to C4 was PEROXIDASE 1 (PRX1, also known as PR9), which was repressed at 12 hpi. Since redox buffering plays a role in plant defenses against pathogens, including viruses [44], we screened annotated DE genes potentially associated with redox buffering. Sixteen DE genes were identified with 11 being responsive to C4, but not to BR, bes1-D and/or bzr1-1D (Table 4; Additional file 3), suggesting C4 expression might disrupt redox homeostasis in a manner unique to C4 and independent of the BRSP.

Table 3 Disease Resistance: Gene ontology enrichment analysis of over-represented, up-regulated genes at 12 h post-induction
Table 4 C4-induced genes associated with redox buffering at 12-hpi

Down-regulated differentially expressed genes uniquely responsive to C4

At 6- and 12-hpi, repressed DE genes were enriched in the GO categories lipid transport and lipid binding (Table 5). The most significantly down-regulated DE genes at 6- and 12-hpi were the lipid transfer proteins (LTP) AZELAIC ACID INDUCED 5 (AZI5), AT5G46890 and AT5G46900. AT5G46890 and AT5G46900 are paralogs of AZI5. A fourth LTP, AT4G12520 (AZI6) was also significantly down-regulated at 12-hpi and represents a recent gene duplication of AZI5, encoding an identical protein, albeit the promoters are differently regulated. Interestingly, the 4 LTPs are differentially responsive to C4 and BR and bzr1-1D. The LTPs are down-regulation by C4, while AZI5 and AZI6 are up-regulated by BR and all 4 of the LTPs are up-regulated in the bzr1-1D gain-of-function mutant (Additional file 3).

Table 5 Gene ontology enrichment analysis of down-regulated genes at 6- and 12-h post induction

At 12-hpi, C4 expression resulted in significant down-regulation of 5 genes associated with cell development with 4 of the DE genes being more specifically associated with trichoblast differentiation (Table 5). Three of the 4 DE genes, COBRA-LIKE PROTEIN 9 (AT5G49270, COBL9), XYLOGLUCAN-SPECIFIC GALACTURONOSYLTRANSFERASE 1 (AT1G63450, GT16) and ADENINE NUCLEOTIDE ALPHA HYDROLASES-LIKE PROTEIN 6 (AT2G03720, MRH6) are regulated by C4, but not by BR, bes1-D or bzr1-1D (Table 5 and Additional file 3). Moreover, at 12-hpi, 6 enriched DE genes associated cell wall homeostasis were down-regulated, with 5 of the DE genes being responsive to C4, but not to BR, bes1-D or bzr1-1D (Table 5 and Additional file 3). Taken together, the results suggest that early transcriptional events down-regulated by C4 impact lipid homeostasis, cell development and cell wall integrity independent of the BRSP.

Differentially expressed miRNAs following C4 induction

MicroRNAs are small non-coding RNAs that are involved in gene regulation of a plethora of biological processes, including development and RNA silencing as an immune response to virus infection [45]. Therefore, a small RNA library analysis was performed to determine early events in regulating miRNA expression following expression of the C4 protein. At 6- and 12- hpi, 11 and 22 DE miRNAs were detected in induced IPC4–28 seedlings, respectively, with 3 being common to both time points (Additional file 5). In control experiments with Sei-0 wild-type seedlings, 1 and 9 DE miRNAs were detected in induced IPC4–28 seedlings, respectively, at 6- and 12-hpi. Five of the 9 DE miRNAs detected at 12-hpi in non-induced IPC4–28 seedlings were also differentially expressed in induced seedling, suggesting a role in developmental regulation.

We used psRNATarget to identify targets of the DE miRNAs. At 12-hpi, miR163 was down-regulated and represents the only DE miRNA that was identified to target transcripts of a C4 regulated DE gene. MiR163 targets PXMT1, a member of the Arabidopsis SABATH methyltransferase gene family that prevents seed germination and modulates root development [46]. Both miR163 and PXMT1 are light inducible and there is an inverse correlation between the levels of miR163 and PXMT1 transcripts after C4 induction. The result suggests that up-regulation of PXMT1 transcript levels, noted above, may, at least in part, be due to the down-regulation of miR163.

Network analysis

A coorelation network was used to identify key regulator points where DE genes following C4 expression might be involved in modulating metabolic processes (Supplementary Fig. S3, Additional file 1). At 12-hpi, a single protein, TAP2 INTERACTING PROTEIN OF 41 kDa-like (TIP41-like) protein (AT3G54000) was found to act as a hub connecting two large nodes. TIP41 is the orthologue of mammalian and yeast proteins that function in the Target-of-rapamycin (TOR) pathway, which modifies cell growth in response to nutrient status and environmental conditions [47]. In Arabidopsis, TIP41 is constitutively expressed in vascular tissue and has been shown to bind and regulate SERINE/THREONINE PROTEIN PHOSPHATASE 2A (PP2A) activity, a phosphatase that plays a role in numerous signaling pathways regulating growth and development and stress responses [47].

Discussion

The BCTV C4 protein is a pathogenic determinate that induces extensive development abnormalities, including oncogenesis. In this study, we looked at early patterns of gene expression following induction of the C4 protein to identify initial events in pathogenesis. Since C4 targets AtSKs [9], we initially assumed that C4-induced gene expression would mimic host-gene expression changes observed in the bes1-D and bzr1-1D gain-of-function mutants. However, transcriptional changes following induction of the C4 protein only partially mimicked gene expression responses documented for bes1-D and bzr1-1D (24, 25; Additional file 3). Based on the results we present a working model for early transcriptional changes following C4 expression (Fig. 5). C4 expression results in BRSP-dependent alterations, presumably through the interaction of C4 with AtSKs. However, the results also underscore the importance of C4 altering the Arabidopsis transcriptome by one or more BRSP-independent mechanisms.

Fig. 5
figure 5

Model showing curtovirus C4 regulated BR-signaling pathway dependent (BRSP-dependent) and BRSP-independent functions at 12-hpi. In the absence of BR, BES1 and BZR1 are phosphorylated by AtSKs, inactivating the transcription factors. In the presence of BR or the BCTV C4 protein, the AtSKs are inactivated, which results in the dephosphorylation and activation of BES1 and BZR1. Activated BES1 and BZR1 locate to the nucleus and control expression of BR-biosynthetic genes and BR-responsive genes. Under constant C4 presence, the result is the disruption of hormone homeostasis. C4 also regulates expression of genes not directly regulated by the BR-signaling pathway (BRSP-independent), having early effects on regulatory pathways affecting cellular processes indicated in blue. The image in this figure was generated by the corresponding author

The expression of DE genes unique to C4 and not regulated by BR, bes1-D and bzr1-1D could be explained in three ways. First, the C4 protein induces DE genes through interacts with host proteins not involved in the BRSP. The BCTV C4 protein has been shown to interact with CLV1 [29], a protein involved in meristem cell proliferation and differentiation. This BCTV C4-host protein interaction, as well as ones that likely have not been identified, might be expected to have a significant effect on cellular processes and pathogenesis. Second, DE genes unique to C4 expression might arise from C4 differentially regulating AtSK functions not associated with BES1 and BZR1. AtSK11 and AtSK21 regulate enzymes and transcription factors involved in stomatal, root hair and lateral root development, auxin and ABA signaling, tolerance to salt stress and pattern-triggered immunity [14, 18]. Third, since C4 binds to and inhibits AtSKs, it functions downstream of BR, BRI1 and BAK1 signaling at the PM. Therefore, in the absence of BR, BRI1 accumulates [48] and would be available for BR-independent functions; such as, the interaction of BRI1 and BAK1 with G proteins that regulate sugar-responsive growth and development [49]. In addition, BRI1 controls vascular cell fate in Arabidopsis root through RECEPTOR-LIKE PROTEIN 44 and phytosulfokine signaling [50]. Therefore, regulatory events upstream of AtSK activities would not be directly impacted by C4 and might, at least in part, explain why some BR-responsive genes are not responsive in the gain-of-function mutants [24, 25].

Down regulated DE genes are primarily unique to C4

At 12-hpi, 5 enriched GO categories (lipid transport, cell development, root epidermal cell differentiation, and the child term trichoblast differentiation, as well as, plant cell wall) were identified from the 20 down-regulated DE genes that were uniquely responsive to C4 (Fig. 4, Additional file 3). Four of the down-regulated DE genes are lipid transfer proteins (LTPs), 3 of which have been shown to be up-regulated in the bzr1-1D mutant (Additional file 3). The 4 LTPs, are members of the AZI1 family of LTPs [51]. AZI1 and the closely related paralog, EARLY ARABIDOPSIS ALUMINUM INDUCED 1, have been implicated in signaling as positive regulators of systemic acquired resistance (SAR), induced systemic resistance signaling, MAMP-triggered SAR, lignin biosynthesis, root growth under zinc-limiting conditions, and abiotic (salt and cold) tolerance [52,53,54,55,56,57]. If the DE LTP genes are involved in promoting lipid-based signaling during biotic and abiotic stress, the C4 protein might disrupt this process, such as defense-related signaling, and establish an environment favorable for BCTV infection.

Four down-regulated DE genes were associated with trichoblast differentiation and are involved in root hair development (Table 5). Three of the DE genes, GOBL9, GT16 and MRH6, uniquely respond to C4, but not to BR, bes1-D or bzr1-1D (Additional file 3). COBL9 and GT16, a root hair specific acid xyloglucan that incorporates galacturonic acids residues into root hair cell walls, are required for root hair tip-directed growth [58,59,60]. MRH6 is required for normal root hair initiation and to restrict development to a single root-hair on each root-hair bearing cell [59]. The fourth DE gene, LEUCINE-RICH REPEAT/EXTENSIN 1 (LRX1), encodes a cell wall localized, chimeric leucine rich-repeat/extension protein that regulates cell wall development and/or polar growth and, subsequently, root hair morphogenesis and elongation [61]. BRs are involved in controlling multiple aspects of root growth and development [62]. Therefore, because expression of C4 mimics high levels of BR by inhibiting AtSK function [9], these findings provide mechanistic relevance to previous phenotypic findings that roots are stunted and root hairs and lateral roots are either absent or developmentally impaired in seedlings ectopically expressing the C4 protein [6]. This suggests that the C4 protein modulates unique alterations to cell development that are independent of those modulated by the BRSP.

At 12-hpi, 6 DE genes associated with the cell wall were down-regulated (Table 5). Five of the DE genes [PEROXIDASE 1, 2, and 24 (PRX1, PRX2, PRX24) and PECTIN METHYLTRANSFERASE 24, and 46 (PME24 and PME46)] were uniquely responsive to C4 (Additional file 3). PRX1, PRX2, and PRX24 encode class III peroxidases that are primarily involved in cell wall stiffening or loosening [63]. PRX1 and PRX2 are paralogs with identical coding sequences. PRX2 is involved in stem lignification in Arabidopsis. A PRX2 deficient mutant leads to a significant decrease in total lignin content, altered lignin structure and cell wall thinning in Arabidopsis stems [64]. PRX1 is found primarily in roots and leaves and is likely involved in cell wall lignification. It is notable that GIBBERELLIN 2-OXIDASE 8 (GA2OX8) is up-regulation by C4 and BR (Additional file 3) and that up-regulation of GA2OX8 correlates with decreased lignification in cell walls [65]. While very little is known about the activity of PME24 and PME46, the down-regulation of both PMEs has been associated with cell wall remodeling as has the down-regulation of the DE gene LRX1 [66]. Although not identified in the GO analysis, EXTENSIN 13, a proline-rich cell wall protein, that is uniquely responsive to only C4, was also significantly down-regulated at 12-hpi (Additional file 3). In addition, at 12-hpi, we observed 3 DE genes that are responsive to BR and/or bes1-D and bzr1-1D with roles in cell wall synthesis that were not identified in the GO analysis, XYLOGLUCAN ENDOTRANSGLUCOSYLASE/HYDROLASE 19 (XTH19), PME41 and AT3G10720 (putative member of the PME family). XTH19 has a role in constructing and deconstructing xyloglucans polymers in primary cell walls, while being involved in cell wall construction of growing tissue. XTH19 has also been shown to be involved in cell proliferation of pith tissue during tissue reunion of incised stem [67]. Interestingly, BR has been shown to regulate expression of a plethora of genes involved in cell wall synthesis and remodeling, including genes with a role in protecting cell wall integrity that results from an imbalance of pectin modifications during growth and development [16, 68]. However, at 12-hpi, we did not see regulation of a large number of BR and/or bes1-D and bzr1-1D responsive DE cell-wall associated genes, an expression pattern that might occur at later times post-C4 expression. The results do suggest that C4 expression regulates the expression of genes involved in cell wall remodeling that are independent of BR-regulated cell wall homeostasis. It is tempting to speculate that the effect of C4 on the cell wall could result in cell wall thinning and remodeling, possible a prerequisite for C4-induced cell division and/or regulation of plasmodesmata molecular exclusion limits for cell-to-cell movement of BCTV. It should be noted that the Tobacco mosaic virus (TMV) movement protein was shown to bind a cell wall associated PME, a requirement for movement of the virus from cell to cell through plasmodesmata [69]. More recently a tobacco PME was shown to enhance RNA silencing induced by TMV, suggesting a role for a PME(s) in a virus defense [70].

C4-induced alterations to hormone homeostasis primarily depend on regulation of the BRSP

Plant growth and development are highly regulated processes involving the endogenous signaling of multiple plant hormones in complex interconnected networks in response to varying environmental conditions [71]. Since C4 negatively regulates AtSKs, it is not surprising that C4-induced transcriptional changes result in disruptions of hormone homeostasis at 6-hpi, which is enhanced at 12-hpi. Of the 24 up-regulated DE genes affecting hormone homeostasis at 12-hpi (Table 2), only 4 were not responsive to the BRSP, suggesting C4 effects multiple hormone-related processes by primarily regulating the BRSP.

However, two up-regulated DE genes that could affect hormone homeostasis and are responsive to C4, but not BR, bes1-D or bzr1-1D, are ACS9 and LOG5 (Additional file 3). Both directly affect hormone synthesis and levels and could enhance the C4-directed BRSP disruption of hormone homeostasis. ACS9 was one of three ethylene biosynthesis enzymes (ACS4, 5 and 9) highly enriched as 6- and 12-hpi. Ethylene regulates a number of developmental and stress responses, including seed germination, fruit ripening, senescence, cell elongation and pathogen defense [72]. ACS4, 5 and 9 are closely related type II ACS proteins that are regulated in a ubiquitin-dependent manner and may negatively regulator plant growth in light [73]. BR and cytokinin regulate type II ACS proteins by stabilize the proteins [74]. While the relevance of C4 up-regulating ACS genes is not clear, high levels of ethylene might contribute to the stunting of primary and lateral roots observed in transgenic Arabidopsis expressing C4 [6]. Many geminiviruses, such as BCTV, are limited to the phloem and replication is restricted to the vascular tissue. Therefore, it is interesting that ethylene promotes cell division during vasculature development in Arabidopsis stems [75], a process that ACS9 might take part in or enhance.

LOG5 is one of a 9-member family of cytokinin-activating enzymes in the direct-activation pathway in Arabidopsis that converts inactive cytokinin nucleotides to the biologically active free-base forms. Therefore, the up-regulation of LOG5 at 12-hpi would be expected to have a direct effect on increasing levels of cytokinin. Cytokinin promotes cell division, cell expansion and differentiation, apical dominance and leaf senescence. LOG5 is expressed in vascular tissue of cotyledons, immature leaves, axillary buds, ovules and roots of Arabidopsis [76]. Therefore, cytokinin biosynthesis could be regulated in a vascular-specific, C4-dependent manner during infection, which might be expected to exacerbate the C4-induced BRSP alterations on hormone homeostasis in vascular tissue.

Defense and redox homeostasis

The primary virus defense responses in plants are thought to be through RNA silencing and resistance (R) genes. In contrast to begomovirus C4/AC4 proteins, there is no evidence that the C4 proteins of curtoviruses are suppressors of RNA silencing and we saw very little evidence that the transcriptional response to C4 was regulated by miRNA activity as early as 12-hpi. However, we cannot exclude the possibility of the DE miRNAs exerting their action through translational repression [77, 78]. Also, begomovirus C4 proteins interact with BAM1 and BAM2 and negatively regulate their role in RNAi spread [30, 31], a function we cannot rule out for BCTV C4. Expression of C4 did up-regulate 14 DE genes related to fungal, bacterial, and oomycete defense (Table 3), but the majority are also responsive to BR and/or bes1-D and bzr1-1D and may have roles independent of a defense response. Alternatively, extended expression of these pathogen-related genes might benefit BCTV and increase host susceptibility. However, the results do suggest that C4 may disrupting redox homeostasis independent of the BRSP, which might play a role in defense functions [44]. Consistent with C4 modulating redox potential, AtSK11, which BCTV C4 binds [9], phosphorylates and activates glucose-6-phophate dehydrogenase (G6PD). G6PD is required for maintaining cellular redox homeostasis. G6PD activity is reduced in an AtSK11 loss-of-function mutant, resulting in increased levels of reactive-oxygen species and an imbalance in redox potential [79]. Therefore, C4 binding to AtSK11 would be expected to disrupt redox homeostasis.

Questions on C4-induced hyperplasia

We did not detect C4-induced DE cell cycle-related genes at 6- or 12-hpi in 7-day old seedlings. However, we previously showed that 3 mitotic markers [CYCLIN A1;1 (CYCA1;1), B1;4 (CYCB1;4) and CYCLIN-DEPENDENT KINASE B2:2 (CDKB2;2)], that are up-regulated in the G2/M transition and M phases of the cell cycle [80], were significantly up-regulated by 24-hpi in C4-expressing seedlings relative to non-expressing seedlings [6]. CYCA1;1 and CDKB2;2 are responsive to C4, but not to BR, bes1-D or bzr1-1D [24, 25]. Similarly, Parks et al. [7] screened transformed Arabidopsis plants constitutively expressing Beet severe curly top virus (BSCTV)-C4 for a number of cell-cycle related genes. They found CYCB1;1, CYCB1;3, CYCB2;2, CDKA;1 and CDKB1;1 were up-regulated in BSCTV infected tissue and to a much higher level in BSCTV-C4 expressing transgenic plants. CYCB1;1 and CDKB1;1 are responsive to BSCTV-C4, but not to BR, bes1-D or bzr1-1D [24, 25]. In addition, CYCLIN-DEPENDENT KINASE INHIBITOR 1 (an inhibitor of the CYCD2;1/CDKA;1 activity), also uniquely responsive to C4, was much more strongly suppressed in C4 expressing plants than in BSCTV infected tissue. Recently, the C4 protein of TLCYnV was shown to interact with N. benthamiana NbSKη, sequestering the kinase to the PM [8]. This interaction was suggested to impair NbSKη directed degradation of NbCycD1;1, resulting in accumulation of active NbCycD1;1 and abnormal cell division [8]. However, this may not be a mechanism of action of the BCTV C4 protein in Arabidopsis since constitutive overexpression of CYCD1;1 under regulation of the 35S promoter does not result in a C4 phenotype, but promotes endoreduplication in Arabidopsis leaves and hypocotyls [81]. In addition, BR represses CYCD1;1 expression in Arabidopsis [24], although we did not observe C4 mimicking this response as early as 12-hpi.

Conclusions

A requirement of BCTV C4 pathogenesis is that C4 sequesters AtSKs to the PM [9, 27]. Apparently, regulating the BRSP by sequestering AtSKs to PM is not a unique developmental tool for geminiviruses. OCTOPUS, a PM and vascular localized, cell-specific suppressor of BR-mediated signaling has been shown to promote phloem differentiation by sequestering AtSK21 to the PM [17]. Consistent with this, BES1 and BZR1 are key regulators of phloem and xylem cell differentiation in vascular tissue [82]. The two latter points are significant since BCTV replication is limited to phloem tissue. Our results suggest that BCTV C4-induced pathogenesis alters the balance between growth and development by usurping both the BRSP and one or more BR-independent signaling pathways (Fig. 5). We suggest that C4-pathogenesis results from the protein interacting with and affecting the outcome of a number of metabolic and/or signaling pathways.

Methods

Plants and transgene induction

Arabidopsis thaliana ecotype Sei-0 (WT) and transgenic Sei-0 line IPC4–28 seedings [6], were used in all experiments. Sei-0 seeds were obtained from Roger Innis (Indiana University, Bloomington, IN) and were originally obtained from the Arabidopsis Information Service seed bank. Line IPC4–28 expresses the BCTV C4 gene from the regulatory control of a ß-est inducible promoter [6]. For RNA-seq analysis, fifty Sei-0 or IPC4–28 seeds were surface sterilized, place at 4 °C for 2 days, and grown in liquid culture for 7 days at 25 °C with gentle shaking (100 rpm) under continuous light [6]. After 7 days, seedlings were mock-induced or induced with 10 μM ß-est and harvested at 0-, 6- and 12-h post-mock-induction (hpmi) or post-induction (hpi), washed with water, blotted dry, frozen in liquid nitrogen and stored at − 80 °C. Three independent experiments were performed for each Arabidopsis line and each treatment.

RNA extraction, RNA-seq and qRT-PCR

RNA-seq results were obtained from RNA extracted from 3 independent biological replicates. Seedlings grown in liquid culture with or without induction with ß-est were homogenized in 0.75 ml of Trizol Reagent in ZR BashingBead Lysis tubes (2.0 mm beads; Zymo Research, Irvin, CA, U.S.A.) on a Geno/Grinder (SPEX SamplePrep 2010, Metuchen, NJ, U.S.A.) for 60 s. at 1750 rpm. RNA was extracted using a Direct-zol RNA Miniprep kit, which included a DNase I on-column treatment, as described by the manufacturer (Zymo Research, Irvin, CA, U.S.A.). Initial quality control of the RNA samples was performed using an Agilent 2100 Bioanalyzer (Agilent, Santa Clara, CA, U.S.A.) to determine RNA purity, concentration and integrity. Sample concentration was normalized in 25uL of nuclease-free H2O prior to library preparation. Libraries were prepared and sequenced by the Georgia Genomics and Bioinformatics Core Lab team (GGBC). Briefly, libraries were prepared using KAPA’s Stranded mRNA-seq kit (cat. KK4821). During library preparation, mRNA was selected using oligo-dT beads, the RNA was fragmented, and cDNA was generated using random hexamer priming. Single indices and Illumina adapters were ligated to make the sequencing libraries. Quality control of the libraries were performed using a Qubit and Fragment Analyzer to determine library concentration and size distribution. Libraries were multiplexed using an equimolar amount and the concentration of the final pool was assessed using qPCR. The final pool was sequenced on an Illumina NextSeq 500 platform using the paired-end (PE) 75 protocol.

The small RNA (sRNA) libraries were prepared and sequenced by the GGBC. Briefly, sRNA libraries were prepared using the Bioo Scientific NEXTFLEX Small RNA-Seq Kit v3 with the following steps: 3′ and 5′ adapters were selectively ligated to the small RNA molecules, and cDNA was synthesized using a primer specific to the 3′ adapter sequence. Index sequences were added to each library during library PCR. Final libraries were size selected to enrich for small RNA products using the bead ratios described in the kit’s protocol with NEXTFLEX Clean-up Beads. Quality control of the prepared libraries was performed using the Qubit and Fragment Analyzer to determine the concentration and size of the libraries. SRNA libraries were multiplexed using equimolar ratios. The final pool was sequenced on the Illumina NextSeq 500 platform using single-end (SE) 75 protocol.

RNA-seq raw reads were quality filtered and trimmed using Trim Galore (Babraham Bioinformatics – Trim Galore) before mapping to the Arabidopsis reference genome (TAIR10) using the short reads quasi-mapper Salmon [83]. Genes that had a total of 10 reads or less across all samples were filtered out. PCA analysis from the package DESeq2 was used to assess the quality of the samples and replicates. The count matrix was normalized and fitted using the default DESeq2 normalization method. Differential analysis was carried out on the normalized matrix using the contrast experimental design (Fig. 1a) using DESeq2 [84].

SRNA-seq raw reads were quality filtered and trimmed using the filter tool in the UEA small RNA Workbench [85]. This step includes removing rRNA and mRNA degradation products as well as low complexity reads and adapter traces. Alignment of small RNA reads to the Arabidopsis sRNA transcriptome, reads count, normalization and differential expression analyses were done using the tools available in the UEA small RNA workbench. The miRNA-mRNA target analysis was done using the psRNATarget tool [86] using TAIR10 reference. Genes or miRNA that were significantly up-regulated or down-regulated following ß-est induction of the C4 protein in transgenic IPC4–28 seedlings relative to mock-induced control seedlings were designated as differentially expressed (DE) if they had log2 fold changes ≥2.0 with adjusted p-values ≤0.05.

For RT-qPCR analysis, RNA extracted and purified for RNA-seq was used, representing three independent biological replicates. cDNAs were synthesized using SuperScript™ III Reverse Transcriptase (Invitrogen) and primers describe in Additional file 6. PCRs were performed on a QuantStudio 5 Real-Time PCR System (Applied Biosystems) using SYBR™ Green PCR Master Mix (Applied Biosystems). MONENSIN SENSITIVITY 1 (MON1, AT2G28390) was used as an endogenous housekeeping control. A relative standard curve method was used to calculate the fold change of gene expression between ß-est induced and mock induced transgenic IPC4–28 plants. C4-induced transcriptional changes were normalized to gene expression in noninduced seedlings.

Network analysis

The network analysis was conducted in three steps: 1) filtration of low count genes and logarithmic transformation, 2) differential expression analysis, and 3) co-expression network construction. Only genes with read counts in all three time points (0-, 6-, and 12-hpi) were used in this analysis. We used the Bioiconductor ‘org. At.tai.db’ to annotate genes [87]. Differential expression analysis was conducted using DESeq2 [84]. Differentially expressed genes with P-value > 0.05 were removed. Gene-Gene similarity matrix was developed using Pearson correlation and Euclidean distances measures. The similarity matrix was converted to adjacency matrix using power transformation [88]. The adjacency matrix was visualized in the Cytoscape software. Only edges with a minimum of 0.75 coefficient score were permitted in the network. Both edges and nodes were colored to reflect gene-gene correlation coefficiency and connectivity.

Availability of data and materials

The datasets generated and analyzed during the current study are available in the Genbank repository, Accession: PRJNA701411. Access at https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/Traces/study/?acc=PRJNA701411&o=acc_s%3Aa

Abbreviations

ABCG40:

ABC TRANSPORTER G FAMILY 40

ACS:

1-AMINOCYCLOPROPANE-1-CARBOXYLATE SYNTHASE

AtSK:

Arabidopsis thaliana shaggy-like protein kinase

AZI:

AZELAIC ACID INDUCED

BAK1:

BRASSIONSTEROID INSENITIVE 1-ASSOCIATED RECEPTOR KINASE 1

BAM:

BARELY ANY MERISTEM

BCTV:

Beet curly top virus

BES1:

BRI1-EMS SUPPRESSOR 1

BR:

Brassinosteroid

BRI1:

BRASSINOSTEROID INSENSITIVE 1

BRSP:

Brassinosteroid-signaling pathway

BSCTV:

Beet severe curly top virus

BZR1:

BRASSINAZOLE-RESISTANT 1

CDK:

Cyclin dependent kinase

CLV1:

CLAVATA 1

CML41:

Calmodulin-like 41

COBL9:

COBRA-LIKE PROTEIN 9

CYC:

Cyclin

DE:

Differentially expressed

G6PD:

Glucose-6-phosphate dehydrogenase

GA2OX8:

GIBBERELLIN 2-OXIDASE 8

GGBC:

Georgia Genomics and Bioinformatics Core Laboratory

GO:

Gene ontology

GT16:

GALACTURONOSYLTRANSFERASE 1

HIR1:

HYPERSENSITIVE INDUCED REACTION 1

IPC4–28:

Inducible promoter C4–28

LRX1:

LEUCINE-RICH REPEAT/EXTENSIN 1

LOG5:

LONELY GUY 5

LTP:

Lipid transfer protein

MRH6:

ADENINE NUCLEOTIDE ALPHA HYDROLASE-LIKE PROTEIN 6

MON1:

MONENSIN SENSITIVITY 1

NbCYCD1:1:

Nicotiana benthamiana cyclin D1.1

NbSKη:

Nicotiana benthamiana shaggy-like protein kinase eta

PCA:

Principle component analysis

PM:

Plasma membrane

PME:

PECTIN METHYLTRANSFERASE

PP2A:

SERINE/THREONINE PROTEIN PHOSPHATASE 2A

PRX:

Peroxidase

RT-qPCR:

Reverse transcription quantitative polymerase chain reaction

PXMT1:

PARAXANTHINE METHYLTRANSFERASE 1

SAR:

Systemic acquired resistance

TIP41:

TAP2 interacting protein of 41

TMV:

Tobacco mosaic virus

TLCYnV:

Tomato leaf curl Yunnan virus

TUB2:

TUBULIN β CHAIN

XTH19:

XYLOGLUCAN ENDOTRANSGLUCOSYLASE/HYDROLASE 19

WT:

Wild type

References

  1. Leisner SM, Schoelz JE. Joining the crowd: integrating plant virus proteins into the larger world of pathogen effectors. Annu Rev Phytopathol. 2018;56:89–100.

    Article  CAS  PubMed  Google Scholar 

  2. Fondong VN. The ever-expanding role of C4/AC4 in geminivirus infection: punching above its weight? Mol Plant. 2018;12:145–7.

    Article  PubMed  CAS  Google Scholar 

  3. Rojas MR, Hagen C, Lucas WJ, Gilbertson RL. Exploiting chinks in the plant’s armor; evolution and emergence of geminiviruses. Annu Rev Phytopathol. 2005;43:360–94.

    Article  CAS  Google Scholar 

  4. Latham JR, Saunders K, Pinner MS, Stanley J. Induction of plant cell division by beet curly top virus gene C4. Plant J. 1997;11:1273–83.

    Article  CAS  Google Scholar 

  5. Lai J, Chen H, Teng K, Zhao Q, Li Y, Lian L, et al. RKP, a RING finger E3 ligase induced by BSCTV C4 protein, affects geminivirus infection by regulation of the plant cell cycle. Plant J. 2009;74:905–17.

    Article  CAS  Google Scholar 

  6. Mills-Lujan K, Deom CM. Geminivirus C4 protein alters Arabidopsis development. Protoplasma. 2010;239:95–110.

    Article  CAS  PubMed  Google Scholar 

  7. Park J, Hwang H-S, Buckley KJ, Park J-B, Auh C-K, Kim D-G, et al. C4 protein of beet severe curly top virus is a pathomorphogenetic factor in Arabidopsis. Plant Cell Rep. 2010;29:1377–89.

    Article  CAS  PubMed  Google Scholar 

  8. Mei Y, Yang X, Huang C, Zhang X, Zhou X. Tomato leaf curl Yunnan virus-encoded C4 induces cell division through enhancing stability of Cyclin D 1.1 via impairing NbSKη-mediated phosphorylation in Nicotiana benthamiana. PLoS Pathog. 2018;14:e1006789.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  9. Mills-Lujan K, Andrews DL, Chou C, Deom CM. The roles of phosphorylation and SHAGGY-like kinases in Geminivirus C4 protein induced hyperplasia. PLoS One. 2015;10:e0122356.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  10. Bi H, Fan W, Zhang P. C4 protein of Sweet Potato Leaf Curl Virus regulates brassinosteroid signaling pathway through interaction with AtBIN2 and affects male sterility in Arabidopsis. Front Plant Sci. 2017;8:1689.

    Article  PubMed  PubMed Central  Google Scholar 

  11. Kim TW, Guan S, Sun Y, Deng Z, Tang W, Shang JX, et al. Brassinosteroid signal transduction from cell-surface receptor kinases to nuclear transcription factors. Nat Cell Biol. 2009;11:1254–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Rozhon W, Mayerhofer J, Petutschnig E, Fujioka S, Jonak C. ASKθ, a group-III Arabidopsis GSK3, functions in the brassinosteroid signaling pathway. Plant J. 2010;62:215–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Saidi Y, Hearn TJ, Coates JC. Function and evolution of ‘green’ GSK3/shaggy-like kinases. Trends Plant Sci. 2012;17:39–46.

    Article  CAS  PubMed  Google Scholar 

  14. Youn J-H, Kim T-W. Functional insights of plant GSK-3-like kinases: multi-taskers in diverse cellular signal transduction pathways. Mol Plant. 2015;8:552–65.

    Article  CAS  PubMed  Google Scholar 

  15. Planas-Riverola A, Gupta A, Betegon-Putze I, Bosch N, Ibanes M, Cano-Delgado AI. Brassinosteroid signaling in plant development and adaption to stress. Development. 2019;146:dev151894.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  16. Zhu JY, Sae-Seaw J, Wang ZY. Brassinosteroid signaling. Development. 2013;140:1615–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Anne P, Azzopardi M, Gissot L, Beaubiat S, Hematy K, Palauqui J-C. OCTOPUS negatively regulates BIN2 to control phloem differentiation in Arabidopsis thaliana. Curr Biol. 2015;25:2585–90.

    Article  CAS  Google Scholar 

  18. Stampfl H, Fritz M, Santo SD, Jonal C. The GSK3/shaggy-like kinase ASKα contributes to pattern-triggered immunity. Plant Physiol. 2016;171:1366–77.

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Houbaert A, Ahang C, Tiwari M, Wang K, de Marcos SA, Savatin DV, et al. POLAR-guided signaling complex assembly and localization drive asymmetric cell division. Nature. 2018;563:574–8.

    Article  CAS  PubMed  Google Scholar 

  20. He JX, Gendron JM, Yang Y, Li J, Wang ZY. The GSK3-like kinase BIN2 phosphorylates and destabilizes BZR1, a positive regulator of the brassinosteroid signaling pathway in Arabidopsis. Proc Natl Acad Sci U S A. 2002;99:10185–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Wang ZY, Nakano T, Gendron J, He J, Chen M, Vafeados D, et al. Nuclear-localized BZR1 mediates brassinosteroid-induced growth and feedback suppression of brassinosteroid biosynthesis. Dev Cell. 2002;2:505–13.

    Article  CAS  PubMed  Google Scholar 

  22. Yin Y, Wang ZY, Mora-Garcia S, Li J, Yoshida S, Asami T, et al. BES1 accumulates in the nucleus in response to brassinosteroids to regulate gene expression and promote stem elongation. Cell. 2002;109:181–91.

    Article  CAS  PubMed  Google Scholar 

  23. Zhao J, Peng P, Schmitz J, Decker AD, Tax FE, Li J. Two putative BIN2 substrates are nuclear components of brassinosteroid signaling. Plant Physiol. 2002;130:1221–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Sun Y, Fan X-Y, Cao D-M, He K, Tang W, Zhu J-Y, et al. Integration of brassinosteroid signal transduction with the transcription network for plant regulation in Arabidopsis. Dev Cell. 2010;19:765–77.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Yu X, Li L, Zola J, Aluru M, Ye H, Foudree A, et al. A brassinosteroid transcriptional network revealed by genome-wide identification of BES1 target genes in Arabidopsis thaliana. Plant J. 2011;65:634–46.

    Article  CAS  PubMed  Google Scholar 

  26. Li Q-F, Lu J, Yu J-W, Zhang C-Q, He J-X, Lui Q-Q. The brassinosteroid-regulated transcription factors BZR1/BES1 function as a coordinator in multisignal-regulated plant growth. Biochim Biophys Acta Gene Regul Mech. 1861;2018:561–71.

    Google Scholar 

  27. Piroux N, Saunders K, Page A, Stanley J. Geminivirus pathogenicity protein C4 interacts with Arabidopsis thaliana shaggy-related protein kinase AtSKeta, a component of the brassinosteroid signalling pathway. Virology. 2007;362:428–40.

    Article  CAS  PubMed  Google Scholar 

  28. Dogra SC, Rezaian MA, Randles JW. A novel shaggy-like kinase interacts with the Tomato leaf curl virus pathogenicity determinant C4 protein. Plant Mol Biol. 2009;71:25–38.

    Article  CAS  PubMed  Google Scholar 

  29. Li H, Zeng R, Chen Z, Liu X, Cao Z, Xie Q, et al. S-acylation of a geminivirus C4 protein is essential for regulating the CLAVATA pathway in symptom determination. J Exp Bot. 2018;69:4459–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Rosas-Diaz T, Zhanga D, Fana P, Wanga L, Dinga X, Jianga Y, et al. A virus-targeted plant receptor-like kinase promotes cell-to-cell spread of RNAi. Proc Natl Acad Sci U S A. 2018;115:1388–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Carluccio AV, Prigigallo MI, Rosas-Diaz T, Lozano-Duran R, Stavolone L. S-acylation mediates Mungbean yellow mosaic virus AC4 localization to the plasma membrane and in turns gene silencing suppression. PLoS Pathog. 2018;14:e1007207.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  32. Hazak O, Hardtke CS. CLAVATA 1-type receptors in plant development. J Exp Bot. 2016;67:4827–33.

    Article  CAS  PubMed  Google Scholar 

  33. Ismayil A, Haxim Y, Wang Y, Li H, Qian L, Han T, et al. Cotton Leaf Curl Multan virus C4 protein suppresses both transcriptional and post-transcriptional gene silencing by interacting with SAM synthetase. PLoS Pathog. 2018;14:e1007282.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  34. Vinutha T, Kumar G, Garg V, Canto T, Palukaitis P, Ramesh SV, et al. Tomato geminivirus encoded RNAi suppressor protein, AC4 interacts with host AGO4 and precludes viral DNA methylation. Gene. 2018;678:184–95.

    Article  CAS  PubMed  Google Scholar 

  35. Mei Y, Ma Z, Wang Y, Zhou X. Geminivirus C4 antagonizes the HIR1-mediated hypersensitive response by inhibiting the HIR1 self-interaction and promoting degradation of the protein. New Phytol. 2019;225:1311–26.

    Article  PubMed  CAS  Google Scholar 

  36. Ascencio-Ibáñez JT, Sozzani R, Lee TJ, Chu T-M, Wolfinger RD, Cella R, et al. Global analysis of Arabidopsis gene expression uncovers a complex array of changes impacting pathogen response and cell cycle during geminivirus infection. Plant Physiol. 2008;148:436–54.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  37. Gongora-Castillo E, Ibarra-Laclette E, Trejo-Saavedra DL, Rivera-Bustamante RF. Transcriptome analysis of symptomatic and recovered leaves of geminivirus-infected pepper (Capsicum annuum). Virol J. 2012;9:295.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Miozz L, Napoli C, Sardo L, Accotto GP. Transcriptomics of the interaction between the monopartite phloem-limited geminivirus tomato yellow leaf curl Sardinia virus and Solanum lycopersicum highlights a role for plant hormones, autophagy and plant immune system fine tuning during infection. PLoS One. 2014;9:e89951.

    Article  CAS  Google Scholar 

  39. Naqvi RZ, Zaidi SS, Akhtar KP, Strickler S, Woldemariam M, Mishra B, et al. Transcriptomics reveal multiple resistance mechanisms against cotton leaf curl disease in a naturally immune cotton species, Gossypium arboretum. Sci Rep. 2017;7:15880.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Xu B, Cheval C, Laohavisit A, Hocking B, Chiasson D, Olsson TS, et al. A calmodulin-like protein regulates plasmodesmal closure during bacterial immune responses. New Phytol. 2017;215:77–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  41. Rajniak J, Barco B, Clay NK, Sattely ES. A new cyanogenic metabolite in Arabidopsis required for inducible pathogen defense. Nature. 2015;525:376–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Zhao P, Zhang F, Liu D, Imani J, Lange G, Kogel K-H. Matrix metalloproteinases operate redundantly in Arabidopsis immunity against necrotrophic and biotrophic fungal pathogens. PLoS One. 2017;12:e0183577.

    Article  PubMed  PubMed Central  Google Scholar 

  43. De Lorenzo G, Cervone F. Polygalacturonase-inhibiting proteins (PGIPs): their role in specificity and defense against pathogenic fungi. In: Stacey G, Keen NT, editors. Plant-microbe interactions. Boston: Springer; 1997. p. 76–93.

    Chapter  Google Scholar 

  44. Frederickson DE, Loake M, Loake GJ. Redox regulation in plant immune function. Antioxid Redox Signal. 2014;21:1373–88.

    Article  CAS  Google Scholar 

  45. Wang J, Mei J, Ren G. Plant miRNAs: biogenesis, homeostasis, and degradation. Front Plant Sci. 2019;10:360.

    Article  PubMed  PubMed Central  Google Scholar 

  46. Chung PJ, Park BS, Wang H, Liu J, Jang I-C, Chua N-H. Light-inducible MiR163 targets PXMT1 transcripts to promote seed germination and primary root elongation in Arabidopsis. Plant Physiol. 2016;170:1772–82.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Punzo P, Ruggiero A, Grillo S, Batelli G. TIP41 network analysis and mutant phenotypes predict interactins between the TOR and ABA pathways. Plant Signal Behav. 2018;13:e1537698.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  48. Wu G, Wang X, Li X, Kamiya Y, Otegui MS, Chory J. Methylation of a phosphatase specifies dephosphorylation and degradation of activated brassinosteroid receptors. Sci Signal. 2011;4:ra29.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  49. Peng Y, Chen L, Li S, Zhang Y, Xu R, Liu Z, et al. BRI1 and BAK1 interact with G proteins and regulate sugar-responsive growth and development in Arabidopsis. Nature Comm. 2018;9:1522.

    Article  CAS  Google Scholar 

  50. Holzwart E, Huerta AI, Glockner N, Gomez BG, Wanke F, Augustin S, et al. BRI1 controls vascular cell fate in the Arabidopsis root through RLP44 and phytosulfokine signaling. Proc Natl Acad Sci U S A. 2018;115:11838–43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. José-Estanyol M, Gomis-Rüth FX, Puigdmenech P. The eight-cysteine motif, a versatile structure in plant proteins. Plant Physiol Biochem. 2004;42:355–65.

    Article  PubMed  CAS  Google Scholar 

  52. Yu K, Soares JM, Mandal MK, Wang C, Chanda B, Gifford AN, et al. A feedback regulatory loop between G3P and lipid transfer proteins DIR1 and AZI1 mediates azelaic-acid-induced systemic immunity. Cell Rep. 2013;3:1266–78.

    Article  CAS  PubMed  Google Scholar 

  53. Pitzschke A, Datta S, Persak H. Salt stress in Arabidopsis: lipid transfer protein AZI1 and its control by mitogen-activated protein kinase MPK3. Mol Plant. 2014;4:722–38.

    Article  CAS  Google Scholar 

  54. Cecchini NM, Steffes K, Schappi MR, Gifford AN, Greenberg JT. Arabidopsis AZI1 family proteins mediate signal mobilization for systemic defense priming. Nat Commun. 2015;6:7658.

    Article  PubMed  Google Scholar 

  55. Gao H, Wang X-Y, Han Y-Y, Du G-L, Fewng H, Xu Z-Q. Accumulation of the azelaic acid-induced protein AZI1 affects lignin synthesis and deposition in Arabidopsis thaliana. Plant Growth Regul. 2015a;75:317–30.

    Article  CAS  Google Scholar 

  56. Gao Q-M, Zhu S, Kachroo P, Kachroo A. Signal regulators of systemic acquired resistance. Front Plant Sci. 2015b;6:228.

    PubMed  PubMed Central  Google Scholar 

  57. Bouain N, Satbhai SB, Korte A, Saenchai C, Desbrosses G, Berthomieu P, et al. Natural allelic variation of the AZI1 gene controls root growth under zinc-limiting conditions. PLoS Genet. 2018;14:e1007304.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  58. Parker JS, Cavell AC, Dolan L, Roberts K, Grierson CS. Genetic interactions during root hair morphogenesis in Arabidopsis. Plant Cell. 2000;12:1961–74.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Jones MA, Raymond MJ, Smirnoff N. Analysis of the root-hair morphogenesis transcriptome reveals the molecular identity of six genes with roles in root-hair development in Arabidopsis. Plant J. 2006;45:83–100.

    Article  CAS  PubMed  Google Scholar 

  60. Peña MJ, Kong Y, York WS, O’Neill MA. A galacturonic acid-containing xyloglucan is involved in Arabidopsis root hair tip growth. Plant Cell. 2012;24:4511–24.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  61. Baumberger N, Ringli C, Keller B. The chimeric leucine-rich repeat/extension cell wall protein LRX1 is required for root hair morphogenesis in Arabidopsis thaliana. Genes Dev. 2001;15:1128–39.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Wei Z, Li J. (2016). Brassinosteroids regulate root growth, development, and symbiosis. Mol Plant. 2016;9:86–100.

    Article  CAS  PubMed  Google Scholar 

  63. Francoz E, Ranosha P, Nguyen-Kim H, Jamet E, Burlat V, Dunand C. Roles of cell wall peroxidases in plant development. Phytochem. 2015;112:15–21.

    Article  CAS  Google Scholar 

  64. Shigeto J, Kiyonaga Y, Fujita K, Kondo R, Tsutsumi Y. Putative cationic cell-wall-bound peroxidase homologues in Arabidopsis, AtPrx2, AtPrx25, and AtPrx71, are involved in lignification. J Agric Food Chem. 2013;61:3781–8.

    Article  CAS  PubMed  Google Scholar 

  65. Zhao X-Y, Zhu D-F, Zhou B, Peng W-S, Lin J-Z, Huang X-Q, et al. Over-expression of the AtGA2ox8 gene decreases the biomass accumulation and lignification in rapeseed (Brassica napus L.). J Zhejiang Univ Sci B. 2010;11:471–81.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  66. Guenin S, Hardouin J, Paynel F, Muller K, Mongelard G, Driouich A, et al. AtPME3, a ubiquitous cell wall pectin methytransferase of Arabidopsis thaliana, alters the metabolism of cruciferin seed storage proteins during post-germinative growth of seedlings. J Exp Bot. 2017;68:1083–95.

    Article  CAS  PubMed  Google Scholar 

  67. Pitaksaringkarn W, Matsuoka K, Asahina M, Miura K, Sage-Ono M, Yokoyama R, Nishitani K, Ishii T, Iwai H, Satoh S. XTH20 and XTH19 regulated by ANAC071 under auxin flow are involved in cell proliferation in incised Arabidopsis inflorescence stems. Plant J. 2014;80:604–14.

    Article  CAS  PubMed  Google Scholar 

  68. Wolf S, Mravec J, Greiner S, Mouille G, Hofte H. Plant cell wall homeostasis is mediated by brassinosteroid feedback signaling. Curr Biol. 2012;22:1732–7.

    Article  CAS  PubMed  Google Scholar 

  69. Chen M-H, Sheng J, Hind G, Handa AK, Citovsky V. (2000). Interaction between the tobacco mosaic virus movement protein and host cell pectin methytransferases is required for viral cell-to-cell movement. EMBO J. 2000;19:913–20.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Dorokhov YL, Frolova OY, Skurat EV, Ivanov PA, Gasanova TV, Sheveleva AA, et al. A novel function for a ubiquitous plant enzyme pectin methyltransferase: the enhancer of RNA silencing. FEBS Lett. 2006;580:3872–8.

    Article  CAS  PubMed  Google Scholar 

  71. Murphy A. Hormone crosstalk in plants. J Exp Bot. 2015;66:4853–4.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Dubois M, van der Broeck L, Inze D. The pivotal role of ethylene in plant growth. Trends Plant Sci. 2018;23:311–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  73. Tsushisaka A, Yu G, Jin H, Alonso JM, Ecker JR, Zhang X, et al. A combinatorial interplay among the 1-aminocyclopropane-1-carboxylate isoforms regulates ethylene biosynthesis in Arabidopsis thaliana. Genetics. 2009;183:979–1003.

    Article  CAS  Google Scholar 

  74. Hansen M, Chae HS, Kieber JJ. Regulation of ACS protein stability by cytokinin and brassinosteroid. Plant J. 2009;57:606–14.

    Article  CAS  PubMed  Google Scholar 

  75. Etchells JP, Provost CM, Turner SR. Plant vascular cell division is maintained by and interaction between PXY and ethylene signalling. PLoS Genet. 2012;8:e1002997.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  76. Kuroha T, Tokunaga H, Kojima M, Ueda N, Ishida T, Nagawa S, et al. Functional analyses of LONELY GUY cytokinin-activating enzymes reveal the importance of the direct activation pathway in Arabidopsis. Plant Cell. 2009;21:3152–69.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  77. Ma X, Cao X, Mo B, Chen X. Trip to ER. RNA Biol. 2013;10:1586–92.

    Article  PubMed  PubMed Central  Google Scholar 

  78. Iwakawa HO, Tomari Y. The functions of microRNAs: mRNA decay and translational repression. Trends Cell Biol. 2015;25:651–65.

    Article  CAS  PubMed  Google Scholar 

  79. Santo SD, Stampfi H, Krasensky J, Kempa S, Gibon Y, Petutschnig E, et al. Stress-induced GSK3 regulates the redox stress response by phosphorylating glucose-6-phosphate dehydrogenase in Arabidopsis. Plant Cell. 2012;24:3380–92.

    Article  CAS  Google Scholar 

  80. Menges M, de Jager SM, Gruissem W, Murray JAH. Global analysis of the core cell cycle regulators of Arabidopsis identifies novel genes, reveals multiple and highly specific profiles of expression and provides a coherent model for plant cell cycle control. Plant J. 2005;41:546–66.

    Article  CAS  PubMed  Google Scholar 

  81. Zhang G, Zhao H, Zhang C, Li X, Lyu Y, Qi D, et al. TCP7 functions redundantly with several class I TCPs and regulates endoreplication in Arabidopsis. J Integr Plant Biol. 2019;61:1151–70.

    Article  CAS  PubMed  Google Scholar 

  82. Saito M, Kondo Y, Fukuda H. BES1 and BZR1 redundantly promote phloem and xylem differentiation. Plant Cell Physiol. 2018;59:590–600.

    Article  CAS  PubMed  Google Scholar 

  83. Patro R, Duggal G, Love MI, Irizarry RA, Kingsford C. Salmon provides fast and bias-aware quantification of transcript expression. Nat Methods. 2017;14:417–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Love MI, Huber W, Anders S. Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 2014;15:550.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  85. Moxon S, Schwach F, Dalmay T, MacLean D, Studholme DJ, Moulton V. A toolkit for analysing large-scale plant small RNA datasets. Bioinformatics. 2008;24:2252–3.

    Article  CAS  PubMed  Google Scholar 

  86. Dai X, Zhuang Z, Zhao PX. psRNATarget: a plant small RNA target analysis server (2017 release). Nucleic Acids Res. 2018;46:W49–54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Carlson M. (2019). org. At.tair.db: Genome wide annotation for Arabidopsis. R package version 3.8.2. doi: https://0-doi-org.brum.beds.ac.uk/10.18129/B9.bioc.org.At.tair.db

  88. Zhang B, Horvath S. A general framework for weighted gene co-expression network analysis. Stat Appl Genet Mol Biol. 2005;4:article 17.

Download references

Acknowledgments

The authors would like to thank Dr. Brian Kvitko for critically reviewing the manuscript, Kisha Shelton for assistance on figures, and Zengyan Wang for running the network analysis. We also thank the Georgia Genomics and Bioinformatics Core (GGBC) for assessing the quality of RNA samples, building the sequencing libraries and doing the sequencing.

Funding

This research was supported by the Georgia Agricultural Experiment Station. The funder had no role in the design and collection, analysis and interpretation of data or in writing the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

CMD conceived the study, performed the experiments and organized the manuscript. MSA lead the sequencing and performed the bioinformatics data analysis. MSA and LY analyzed the miRNA. LY performed the RT-qPCR analysis. CMD and MSA wrote the manuscript. LY edited the manuscript. All authors have read and approved the manuscript.

Corresponding author

Correspondence to Carl Michael Deom.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: Supplementary Figure S1.

A histogram comparing the counts of quality-trimmed reads and mapped reads for each biological replicate (Rep1, Rep2, and Rep3) from transgenic (Trans) or wild-type (WT) seedlings at 0-, 6-, and 12-hours post-induction (hpi) or hours post-mock induction (hpmi) as well as the overall mapping rate. I, induced. NI, not induced. Supplementary Figure S2. A principal component analysis (PCA) of the replicates using the read count in the gene-condition count matrix. The PCA shows clustering of replicates and the variance/distance between 6- and 12-h post-induction (hpi) of C4Trans_I replicates and replicates of all other conditions. Trans, transgenic seedlings. WT, wild-type seedlings. I, induced. NI, not induced. Hpmi, hours post-mock induction. Supplementary Figure S3. Co-expression network of DE genes. Differentially expressed genes at 12 h post-induction. Nodes are colored based the node degree of connectivity and edges are colored based on correlation coefficient values.

Additional file 2: Table S1.

List of significant differentially expressed genes at 6- and 12-hpi and list of average C4 mRNA reads/sample.

Additional file 3: Table S2.

Responsiveness of C4 differentially expressed genes at 12-hpi relative to brassinosteroid and bes1-D and bzr1-1D gain-of-function mutants.

Additional file 4: Table S3.

Gene ontology (GO) terms indicating over-represented, up-regulated genes in hormone homeostasis at 6 h post-induction.

Additional file 5: Table S4.

Significantly expressed miRNAs.

Additional file 6: Table S5.

Primers used for RT-qPCR.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Deom, C.M., Alabady, M.S. & Yang, L. Early transcriptome changes induced by the Geminivirus C4 oncoprotein: setting the stage for oncogenesis. BMC Genomics 22, 147 (2021). https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-021-07455-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://0-doi-org.brum.beds.ac.uk/10.1186/s12864-021-07455-y

Keywords