Next Article in Journal
Comparative Effects of Different Light Sources on the Production of Key Secondary Metabolites in Plants In Vitro Cultures
Previous Article in Journal
Trait-Environment Relationships Reveal the Success of Alien Plants Invasiveness in an Urbanized Landscape
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Seeing the Forest through the (Phylogenetic) Trees: Functional Characterisation of Grapevine Terpene Synthase (VviTPS) Paralogues and Orthologues

Department of Viticulture and Oenology, Faculty of AgriSciences, South African Grape and Wine Research Institute, Stellenbosch University, Stellenbosch 7600, South Africa
*
Author to whom correspondence should be addressed.
Plants 2021, 10(8), 1520; https://doi.org/10.3390/plants10081520
Submission received: 22 June 2021 / Revised: 9 July 2021 / Accepted: 10 July 2021 / Published: 26 July 2021
(This article belongs to the Special Issue The Roles of Bioinformatics on Plant Genomics Research)

Abstract

:
Gene families involved in specialised metabolism play a key role in a myriad of ecophysiological and biochemical functions. The Vitis vinifera sesquiterpene synthases represent the largest subfamily of grapevine terpene synthase (VviTPS) genes and are important volatile metabolites for wine flavour and aroma, as well as ecophysiological interactions. The functional characterisation of VviTPS genes is complicated by a reliance on a single reference genome that greatly underrepresents this large gene family, exacerbated by extensive duplications and paralogy. The recent release of multiple phased diploid grapevine genomes, as well as extensive whole-genome resequencing efforts, provide a wealth of new sequence information that can be utilised to overcome the limitations of the reference genome. A large cluster of sesquiterpene synthases, localised to chromosome 18, was explored by means of comparative sequence analyses using the publicly available grapevine reference genome, three PacBio phased diploid genomes and whole-genome resequencing data from multiple genotypes. Two genes, VviTPS04 and -10, were identified as putative paralogues and/or allelic variants. Subsequent gene isolation from multiple grapevine genotypes and characterisation by means of a heterologous in planta expression and volatile analysis resulted in the identification of genotype-specific structural variations and polymorphisms that impact the gene function. These results present novel insight into how grapevine domestication likely shaped the VviTPS landscape to result in genotype-specific functions.

Graphical Abstract

1. Introduction

The Vitis vinifera reference genome has been an invaluable resource for the identification of genes involved in metabolic processes of agronomic interest [1]. The genetic basis for various traits associated with flavour and aroma have increasingly been elucidated since the genome became available, as recently reviewed by reference [2]. A major contributor to flavour and aroma profiles are the terpenes, a chemically diverse class of metabolites, mainly associated with floral and Muscat aromas (for example, linalool, geraniol, nerol, α-terpineol, and hotrienol) or a pepper aroma (for example, rotundone) inferred by mono -and sesquiterpenes, respectively [2,3,4,5,6]. The near-homozygous reference genome PN40024 contains 152 V. vinifera terpene synthase (VviTPS)-like loci, of which 69 are predicted to be functional [1,7]. Thirty of these gene models were subsequently linked to a functional enzyme but were isolated from various genotypes [7].
The VviTPS family annotated on the PN40024 reference genome largely encodes for genes involved in mono- (C10) and sesquiterpene (C15) biosynthesis. These metabolites are biosynthesised via two compartmentalised pathways. The cytosolic mevalonate (MVA) and plastidial 2-C-methyl-D-erythritol 4-phosphate (MEP) pathways both result in the C5 terpene precursors isopentenyl diphosphate (IPP) and dimethylallyl diphosphate (DMAPP). These isomeric precursors are coupled in a head-to-tail manner by prenyl transferases that result in the mono-TPS substrate geranyl diphosphate (GPP) and sesqui-TPS substrate farnesyl diphosphate (FPP). FPP can be isomerised in a rate-limiting reaction to form the nerolidol diphosphate (NPP) substrate [8,9,10,11]. The array of terpenes produced by a single TPS varies greatly, from a single product to an excess of 40 terpenes [12,13,14,15]. This is due to how the TPS active site interacts with its substrate. Depending on how long the TPS can shield the reaction from being quenched, one will typically see the biosynthesis of acyclic and cyclic terpene structures, with the former usually seen due to premature/early quenching [11,16,17,18]. The added double bond in the C15 FPP/NPP substrates, relative to GPP, allows for more conformational changes to take place, catalysed by a sesqui-TPS, resulting in a more diverse array of chemical structures [14,19,20].
The near-homozygosity (~93%) of the PN40024 reference genome [1,21] does not reflect the structural complexity associated with commercial heterozygous diploid cultivars [22,23]. To address this limitation, recent sequencing efforts employed long-read genome sequencing that allows haplotype-aware genome assemblies [24,25,26]. Comparative analyses of the genomic landscape for the diploid genomes of Cabernet Sauvignon (CS), Chardonnay (CH), or Carménère (CR) [24,25,26] highlighted extensive genotype-specific structural variations (SV) within the VviTPS family [27]. The extensive duplication of VviTPS genes furthermore resulted in numerous genes having similar functions with various mechanisms potentially involved in these duplication events [7,20,28,29]. The physical proximity of related VviTPS subfamily members on a chromosome are indicative of tandem duplication events, whereas remnants of transposable elements suggest that transposon-mediated duplication allowed for the genome-wide movement of VviTPS genes [7]. Segmental duplications that are highly homologous (94% and >10 kb in size) contribute to 17% of the PN40024 genome, resulting in large repetitive genomic regions [30]. Domestication-driven introgression between germplasms has furthermore resulted in highly complex genomes that show greater heterozygosity in the domesticated varieties than the wild parent(s) [31,32,33,34].
Tandem TPS duplications often form multi-gene clusters of varying sizes but show similar catalytic mechanisms. In tomatoes (Solanum lycopersicum), for example, a sesqui-TPS cluster of six genes consists of two pseudogenes and characterised genes with divergent catalytic active sites [35]. In rice (Oryza spp.), the evolution of TPS gene clusters involved in insect defence shows species-specific cluster differences that go through a combination of duplication and subsequent functionalisation events. These gene clusters, furthermore, show species-specific expressions and emission patterns in response to insect infestations [36].
A 690-kb region of PN40024 chr18 contains 44/152 VviTPS-like loci, 20 encoding for putative sesquiterpene synthases (VviTPS01-19 and -30). Eleven of these genes have been functionally characterised [7,20]. The annotation of the VviTPS families for CH, CS, and CR, however, revealed that there are potential structural differences in the genome space for the chr.18 cluster [27]. To date, the relatedness of the VviTPS paralogues in this cluster has not been explored.
Of particular interest is VviTPS10, a gene model linked to two distinct enzyme functions with the VvGwaBer nucleotide sequence being concordant to the gene model [7]. A genotype-specific homologue, VviMATPS10, however, showed various nonsynonymous mutations that resulted in a novel function for the cultivar Muscat of Alexandria [20]. The extent of these sequence differences presented two possibilities: VviTPS10 has two alleles with unique functions or a highly conserved paralogue inferring the second function. VviTPS04 was previously identified as a putative homologue or allelic variant of VviTPS10 using the PN40024 reference genome [1,7,20].
In this study, a cluster of chromosome (chr.) 18 genes, which includes VviTPS10 and -04, were analysed in order to explore the possible allelic and/or duplication differences. The annotated VviTPS landscape of three phased diploid genomes [27] was utilised in combination with the publicly available whole-genome resequencing (WGRS) data [37] to identify the structural differences in the selected genotypes. The candidate gene regions homologous to VviTPS10 were subsequently isolated from multiple V. vinifera genotypes and functionally characterised by means of Agrobacterium-meditated transient gene expression in Nicotiana benthamiana, followed by a volatile terpene analysis using headspace solid-phase microextraction (HS-SPME) and gas chromatography-mass spectrometry (GC-MS). The active sites of isolated paralogues were also compared to identify amino acid residues that could result in possible genotype-specific enzyme functions.

2. Results

2.1. Isolated VviTPS04 and -10-Like Paralogues

The PN40024 genome showed that the forward strand of the VviTPS04 locus had a start codon 27 bp upstream of the start codon for the VviTPS10 locus [1,7]. These two loci had high homology with a total of 28 VviTPS-like loci from the three phased diploid genomes (8 in CH, 9 in CR, and 11 in CS), predicted by Smit et al. (2020). The 28 phased diploid and two PN40024 gDNA models (hereafter referred to as reference sequences) were subsequently used to map short reads from 11 different genotypes that were used in the WGRS effort by reference [37]. This resulted in 112/330 consensus sequences with near-complete coverage, with a further 35 that lacked coverage at the 5′ terminal region, relative to the reference sequence. Only a single reference sequence (CRTPS105) did not result in the recovery of at least a single consensus sequence from the 11 WGRS genotypes selected for this analysis. This resulted in 147 consensus and 30 reference sequences that were subsequently aligned and their phylogenetic relationships assessed, as shown in Figure 1 (a higher-resolution version of the complete tree is available in Supplementary Figure S1). The lowest pairwise identity in this alignment was 89.2%, with the highest being 99.5%, indicating that all sequences in the multiple sequence alignment (MSA) were highly similar. The phylogenetic tree revealed that VviTPS04 and VviTPS10-like sequences were grouped into a distinct clade, coloured in grey in Figure 1. Eight of the sequences in this grey clade lacked coverage at the terminal end. The VviTPS04-like sequences (containing the additional 27 bp at the 5′ terminal end) could be separated further from the VviTPS10-like sequences, forming distinct subclades. The VviTPS10 gDNA-containing subclade had an orthologous gDNA model from CS (CSTPS086) with eight gDNA models predicted from the WGRS data. The subclade containing VviTPS04-like sequences had two near-identical reference sequences (CSTPS051 and CHTPS062) to VviTPS04. Ten gene models in this clade were predicted from the WGRS mapping. Incongruence between the PN40024 VviTPS04 gene model and the gene model predicted for the Pinot noir TA-379 WGRS genotype was observed. The short branch lengths indicated high homology for sequences within the grey clade, despite the bifurcation for the VviTPS04-like and VviTPS10-like subclades. The remaining sequences were grouped into five distinct clades (Supplementary Figure S1). These sequences were near-identical at the 5′-region, with 24 lacking coverage. All 175 sequences in the MSA had a near-identical 3′-region. The high homology in the terminal ends meant that three primers could be designed to target 112/175 sequences: Primer A and B targeted the 27-bp difference at the 5′ region, while a single conserved 3′ primer was designed. These results are shown in Table 1, where groups A and B correspond to the binding of their respective forward primers, whereas group C reflects the sequences that lack coverage at these primer-binding regions. The primer-binding circles in the phylogenetic tree (Figure 1) use the same bin colours used in Table 1. Nucleotide sequences for the reference genes and those predicted by mapping of the WGRS data is available is Supplementary File S1.
It should be noted that the 27-bp difference results in two start codons being in frame, shown in Figure 2C. The primers were validated through PCR amplification using gDNA from nine genotypes, as shown in Figure 2A,B. The expected gDNA band of ±2260 bp was present in all genotypes, with certain genotypes showing a second band at ±1700 bp for primer set B (Figure 2B). A second band was notably absent for CH using either primer combination. For primer set A, a 2000-bp amplicon was observed in PN, with this band also faintly visible in six other genotypes (Figure 2A). Primer set A resulted in a unique second amplicon for CS. These results support the in silico predictions and highlighted the extensive homology between the VviTPS04 and VviTPS10-like loci.

2.2. Isolation and Sequence Comparison of VviTPS04 and -10 Homologs from Nine Grapevine Genotypes

Isolation of the genes from grapevine flower cDNAs resulted in the cloning for six out of nine genotypes using primer set A and seven out of nine genotypes using primer set B. No amplification occurred from WR. CS resulted in two amplified sequences for both primer sets. The cDNA amplicon sizes were, however, different for the genotypes (Figure 3A). Sequencing confirmed that two different forms were isolated and that the 27bp differences were sufficient for the isolation of genotypic variants, as evident from the pairwise sequence alignments in Supplementary File S2. Of the 15 isolated sequences, only six had fl-ORFs, while the remaining nine genotypic variants were rendered non-functional due to premature stop codons, frame shifts, and/or intron retention (Figure 3B). The MSA for Figure 3B is available in Supplementary File S3.

2.3. Transient Expression of fl-ORFs of VviTPS10 in Tobacco

The transient expression in N. benthamiana showed that all fl-ORF VviTPS10s produced volatile sesquiterpenes. Isolate VI-A was the only functional gene, other than MA-B (VviTPS10), which produced (E)-β-farnesene as a single product. The VI genotype was the only cultivar that resulted in a functional gene using both primers, with VI-B producing β-caryophyllene and an unidentified sesquiterpene (Figure 4). The unidentified sesquiterpene had a retention time very near to that of (E)-β-farnesene, although its m/z pattern confirmed that it was not (E)-β-farnesene. No definitive match in either the Wiley 275 or NIST libraries could be found due the signal-to-noise ratio of this peak being below the limit of detection (LOD). All other genotypes produced β-caryophyllene and the unidentified sesquiterpene.

2.4. Sequence–Function Relationships of Isolated Paralogues

Querying the phased diploid VviTPS models [27] for the VviTPS10 paralogues revealed that they are part of a cluster consisting of 21 putative proteins that connect to four functionally characterised proteins (Figure 5C). The amino acid sequence similarity of the active sites to that of the characterised enzymes [7,19,20] showed that the proteins in this cluster are predicted to use NPP as the initial substrate, with all, except VviMATPS10, predicted to have an initial 1,6-cyclisation, as predicted by reference [27]. A subsequent analysis of the associated sequences (Figure 5A) showed that these proteins are encoded by a set of genotype-specific duplicated genes. These duplications form highly connected networks, illustrating the extent of paralogy, with the most extensive network connectivity (i.e., greatest number of putative duplications) seen in CS (Figure 5A). Complete gene models associated with the gene duplicates CSTPS083 and CSTPS090 were the closest paralogues of the functional isolates MA-B, CS-1-B, and CB-B (Figure 4B).
Genomic regions where the duplication clusters (Figure 5A) are located were analysed for gene synteny with the differences in the VviTPS landscape of the genotypes, as shown in Figure 6. VviTPS04 had synteny to two genes located on distinct CS contigs, namely CSTPS051 and CSTPS083. VviTPS10 shared synteny with CSTPS051, while it also showed synteny to CSTPS083. CSTPS083 and CSTPS086 are 0.12-Mb apart with three genes in between, two of which are VviTPS genes, in contrast to VviTPS04 and -10, which are 0.27-Mb apart, with seventeen genes in between them, seven being VviTPS genes. A single CH gene, CHTPS067, had synteny to both VviTPS04 and -10. The gene synteny analysis was therefore in agreement with the protein sequence homology represented in Figure 5A.
CRTPS019 and CRTPS199 were identified as potential gene duplications in CR, with the former not predicted to encode for a functional gene. These two gene models were, furthermore, located on a primary contig and haplotig, respectively, that, unlike the CH and CS paralogues, map to two different chromosomes (chr. 13 and 18). For this reason, we did not include CR for the synteny analysis.
The functional genes PN-B and VI-B had the highest I’ scores with the CRTPS019 gene model (97.31 and 97.08, respectively), while VI-A was linked to CRTPS199 with a score of 98.99 (Figure 5B). The I’ score [38] quantifies the extent of paralogy with phylogenetically similar genes (Figure 4B), linking to distinct proteins based on this score (Figure 5B). This allowed for inference of the protein and gene structures most similar to the isolates.
Although five of the isolates were connected to CH gene models, none of these isolates were predicted functional (Figure 3B). A closer inspection of the active site amino acids of derived protein sequences revealed that VI-A and MA-B, (E)-β-farnesene synthases, had near-identical active sites, with the only difference to MA-B being an amino acid deletion and a single nonsynonymous mutation, shown in Figure 7. Furthermore, pairwise comparisons of the full-length proteins for VI-A, VI-B, and MA-A revealed that these proteins were more similar between cultivars than within a cultivar (Supplementary File S4). The active sites of CB-B and CS-1-B were very similar to that of VvGwaBer, with only two amino acid differences (Figure 7). The heterologous in planta expression (Figure 4A), however, did not result in the synthesis of any compounds similar to those observed in vitro for VvGwaBer [7].

3. Discussion

3.1. Genotype-Specific Structural Variation for Isolated VviTPS Paralogues

Genotypic variations in VviTPS-encoding genes were suggested to have a significant impact on the genetic potential of a given cultivar [20]. By targeting two paralogues in nine genotypes, the extent of the structural differences and sequence variations became evident (Figure 1 and Figure 2). Previous efforts focussed only on the enzyme function, with only two studies reporting on the isolation of new and/or cultivar-specific VviTPS genes [19,39]. For example, two point mutations in VviTPS24 were linked to an altered function in Shiraz [19]. The use of a single reference genome in an “applies-to-all” manner is therefore limiting when trying to explore both sequence and structural variations for VviTPS-encoding genes. It was previously suggested that the genotype-specific landscape of VviTPS-encoding genes is greatly underrepresented in the reference genome when compared to next-generation phased diploid genomes [27]. Comparative analyses of the reference [1,21] and diploid genomes [24,25,26] indicated that genotype-specific SV and/or allelic differences likely resulted in a cluster of VviTPS genes on chr. 18 [27]. The results presented here provided functional validation in multiple genotypes for two paralogues in this cluster, namely VviTPS04 and -10.
Once assembled, the diploid genomes should provide new scaffolds for the mapping and mining of SV and SNVs [33,37,40,41,42,43,44] to comprehensively study the allelic differences that impacted the VviTPS functionality. The WGRS data provided valuable insight into the events that shaped the genotypic variations [37,40]. The generation of consensus sequences using gDNA models predicted from phased diploid genomes is advantageous for studying expanded gene families, especially those involved in specialised metabolism. The surveying of WGRS data for 11 different genotypes resulted in 147 consensus sequences with >89% identity to the VviTPS gDNA models predicted from the phased diploid [24,25,26] and reference genomes [1,21]. Although the VviTPS04 and VviTPS10-like sequences formed a distinct clade (Figure 1), extensive conservation at the terminal ends (Table 1 and Figure 1) presented a challenge when designing primers to isolate these genes. The in silico analysis of the primer-binding regions provided important insight into the terminal sequence conservation of the targeted paralogues. The consensus sequences, furthermore, allowed us to identify a phylogenetic clade that is specific to VviTPS04 and -10. This reduced the number of gDNA models that could likely encode for these two genes. This greatly increased the available resolution in terms of sequence information, where we previously relied largely on the PN40024 reference genome [1,21]. This approach can be applied to identify and cross-reference VviTPS genes of interest in cultivars for which no genome sequence exists. The subtle impact of SNPs, which can render a gene non-functional or alter its catalytic mechanisms, can therefore be explored with a finer focus.
The isolated paralogues had extensive homology at the terminal ends; however, when aligned to the aforementioned gDNA sequences, the isolates showed homology to potentially different genomic regions (Figure 5B). The current lack of assembly for the phased diploid genomes does not allow for definitive chromosome positioning. The advantage of PacBio sequencing does, however, allow for the analysis of large contigs. We previously mapped VviTPS-containing contigs to chromosomes using PN40024 as a reference, which allowed for the extrapolation of chromosome localisation [27]. The genotype-specific expansion of VviTPS04 and VviTPS10-like paralogues (Figure 5A and Figure 6) suggests that they are in a chr. 18 region that consists of multiple recent duplications that went through genotype-specific neo-functionalisation, likely exacerbated by domestication and continuous vegetative propagation [41,45].
The CS regions where these paralogues are found (Figure 6) revealed that neither of the primary contigs had a gene order similar to that of PN40024. Assuming that phased sequencing of the two CS contigs in Figure 6 accurately represents two separate genomic regions, one can conclude that these two VviTPS regions are part of a much larger (~950 kb) segmental duplication. Alternatively, these two contigs could represent heterozygous allelic regions on chr.18. The misidentification of allelic regions is a limitation of FALCON-UNZIP (the haplotype aware sequence assembler of PacBio sequences), where heterozygous alleles are assembled into primary contigs instead of haplotigs. New algorithms (for example, FALCON-PHASE) have, however, been developed to address this [46]. Whether these are heterozygous alleles or segmental duplications does not change the fact that there are distinctly different genomic regions associated with the VviTPS clusters from the different genotypes. The lack of assembled genomes is, however, temporary and will likely be addressed in future.
It was evident that the presence of a transcript does not equate to a functional enzyme, with only 6/15 genes shown to be functional (Figure 3 and Figure 4). Extensive SV between genotypes rendered most isolates non-functional; however, distinct active site forms were identified (Figure 7). The isolation of a second genotype with an active site similar to that of VviMATPS10 [20] provided evidence of a unique active site for (E)-β-farnesene biosynthesis in the grapevine. It is not yet clear which amino acid differences are key to this function. The paralogue sequences therefore provided a biologically relevant sequence reference for possible future studies focussed on modelling the protein structure to identify key functional residues for site-directed mutagenesis.
The sequence differences in the terminal end could be used to isolate two distinct loci; however, the VI and MA variants involved in (E)-β-farnesene biosynthesis were isolated with different primer pairs. From this, it was concluded that the 27-bp difference used to distinguish paralogues had no significant impact on the enzyme function. The N-terminal region of a TPS is generally accepted to play a structural role (i.e., correct folding of the protein), while the C-terminal, which contains the metal-binding motifs, is considered to be the catalytic region. It is evident from the active site alignment (Figure 7) that two distinct catalytic sites are associated with the functional isolates. VvGwaBer results in α-bergamotene as the major product (56%), with minor amounts of nerolidol, (E)-β-farnesene (8%), (Z)-α-farnesene (5%), and an unknown terpene (14%) [7]. The catalytic site of VviMATPS10, producing 100% (E)-β-farnesene, is likely more constrictive and/or provides less protection from a nucleophilic attack, resulting in earlier quenching. The evolutionary origin of this function is yet to be determined and will require further analysis of the chromosome structure.

3.2. The Impact of Domestication on VviTPS Expansion

It is theorised that the diversity in natural terpenes provides a fitness advantage with different models proposed for the origin of this diversity, as reviewed in Pichersky and Raguso (2018) [47]. These models, however, do not account for the influence of domestication within a species; instead, they focus on species-wide evolution. The majority of commercial grapevine cultivars have a shared ancestry and are maintained primarily through vegetative propagation to preserve the desired phenotypic traits [48].
The grapevine genomic structure has been shown to be heavily impacted by somatic variations [45,49]. Of particular interest is the occurrence of chromosome shattering (chromothripsis), where large parts of chromosomes are deleted, resulting in the hemizygous loss of large genomic regions, or incorrect joining of such regions, resulting in clustered rearrangements. This phenomenon was linked to berry colour development, where the hemizygous loss of a genomic region, coupled with specific mutations in regulatory genes of anthocyanin biosynthetic genes, resulted in white grapes [49]. Various studies have quantified the extent of SV and/or SNVs between grapevine genotypes [30,44,50], with a comparison between wine and table grapes showing up to 8% of the genome being affected [51]. The VviTPS family has likely been subjected to multiple genome level events, which include segmental duplications [30] and transposon-mediated duplications, evident by the presence of transposon-like remnants around VviTPS genes [7].
The VviTPS genomic landscape of chr. 18 (Figure 6) suggests that large-scale genome rearrangements [49], compounded by introgression (between species) and admixture (within species) [34,41,52], resulted in heterozygous clustered duplications of VviTPS genes. The paralogue/duplication networks in Figure 5A illustrate how these large-scale genomic events resulted in numerous genes of a similar function being localised in a clustered manner. When one considers these clusters in the context of adaptive introgression, where large genomic regions are introduced from a different species, it is highly likely that geographic expansion resulted in the incorporation of alleles adapted to specific environments [34,53]. These alleles have been subjected to extensive mutational changes due to vegetative propagation, allowing for genotype-specific functionality (Figure 3 and Figure 4). It is therefore highly likely that a combination of duplications, chromothripsis, extensive introgression, and SNV events resulted in genotype-specific VviTPS clusters, with the primary assemblies of the CH and CS genomes (Figure 6) supporting this hypothesis. In this study, two paralogues were explored and, although limited, provided valuable insight into how the genomic landscape can shape a cultivar’s potential to produce terpenes. Zhou et al. (2017) suggested that clonal propagation results in the accumulation of deleterious mutations on the recessive allele, which could explain the SNV observed that rendered nine of the 15 VviTPS genes non-functional. Clonal differences within a single cultivar therefore adds even further complexity to mutational events that impact gene functions [24,54].

3.3. VviTPS04 and -10 Functions: An Ecophysiological Perspective

Continuous vegetative propagation ensures that desired traits are maintained but simultaneously provides a relatively unchanging host for pathogens and pests. In order to maintain fruit quality and production levels, European growers apply between 12 and 30 fungicide and 1–10 pesticide treatments per season, depending on the region where the vines are cultivated [55]. VviTPS04 and -10 paralogues are of particular ecophysiological interest due to their role in the biosynthesis of (E)-β-farnesene and (E)-β-caryophyllene, both known kairomones for the grapevine berry moth Lobesia botrana [56,57,58]. This moth causes widespread production losses for grapevines planted in the Mediterranean region, with recent reports of the polyphagous relative L. vanilana causing damage to grapes in a geographically localised region in the Western Cape of South Africa [59]. Due to the commercial losses caused by the grape berry moth, alternative methods to disrupt the kairomone cues through genetic engineering of grapevines was explored by the overexpression of Aaβ-FS, an (E)-β-farnesene synthase orthologue from Artemisia annua [58,60]. The results from this study were promising, as it decreased moth attraction to grapevines. The isolation and functional characterisation of a grapevine (E)-β-farnesene synthase [20] therefore presents a new target for silencing and/or overexpression studies aimed at disrupting L. botrana kairomones. (E)-β-caryophyllene is, however, a second major grapevine sesquiterpene that, along with (E)-β-farnesene and the homoterpene (E)-4,8-dimethyl-1,3,7-nonatriene (DMNT), form the core volatiles that attract the grapevine berry moths [56,57,58]. Eight genes, including the ones from this study, have so far been linked to (E)-β-caryophyllene biosynthesis in grapevines [7,20], which will make it difficult to create a null background for this sesquiterpene.

4. Conclusions

Genome-wide and gene-specific variations resulted in the expansion of the VviTPS family, resulting in tremendous genetic diversity. The paralogues characterised in this study provided valuable insight into how SV and SNP impact gene functions within and between genotypes. WGRS data coupled with the increasing number of diploid Vitis spp. genome sequences will undoubtedly provide valuable new genetic targets to study VviTPS gene variations. Our integration of WGRS with sequenced diploid genomes allowed for an in-depth analysis of the genotypic differences for loci believed to be duplications. Furthermore, we showed that there are extensive gene synteny differences for the CS and CH genotypes when compared to the VviTPS gene cluster on chr. 18 of the PN40024 reference genome. This provides further evidence for the need to move beyond the use of a single reference genome when studying expanded gene families.
The complex biomechanical aspects of isolated VviTPS genes were shown through functional characterisation of the paralogues. Two distinct active sites were identified with transient heterologous expressions showing different biochemical properties for the genotypic isolates. By utilising genomic and WGRS resources to consider both the extent of paralogy and heterozygosity, it should now be possible to more accurately target VviTPS genes to establish structure–function relationships for specific genotypes. Extending our methodology to other expanded gene families—for example, cytochrome P450s—will allow for the identification of genotype-specific variations and/or novel genes.

5. Materials and Methods

5.1. Identification of Putative Paralogues Using Diploid Grapevine Genomes

VviTPS04 and -10 gene models are highly homologous on the PN40024 genome, positioned 262 kb apart on chr18 [7,21]. VviTPS10 was previously shown to have multiple putative paralogues on the Cabernet Sauvignon (CS) genome [20]. The analysis of the paralogues was extended to the Chardonnay (CH) and Carménère (CR) genomes, and the CS genome was re-evaluated using the respective curated gene models reported by reference [27]. The curated VviTPS04- and -10 genomic sequences [7] were retrieved from FLAGdb++ [61] and queried with BLAST [62] to identify their homologous gDNA sequences in the phased diploid genomes.

5.2. Identification of Putative VviTPS04 and -10 Structural Differences Using Whole-Genome Resequencing Data

Short-read sequencing data of eleven Vitis accessions were retrieved from the Sequence Read Archive (SRA) under BioProject ID PRJNA388292 [37]. SRA data was converted to the FASTQ format using the SRA Toolkit (https://github.com/ncbi/sra-tools (Accessed on 25 June 2020)). Magic-BLAST [63] was used to map short reads to genomic reference sequences. The resulting Sequence Alignment Map (SAM) files were analysed using Geneious Prime v2020.1.2. (Biomatters Ltd., Auckland, New Zealand) The 330 SAM mappings generated were manually assessed to ensure sufficient coverage relative to the reference sequence, with no internal gaps in coverage, followed by the extraction of consensus sequences. The CLUSTAL-Omega algorithm [64,65] was used for multiple sequence alignments (MSA) using Geneious Prime v2020.1.2. (Biomatters Ltd., Auckland, New Zealand. The subsequent phylogenetic tree was constructed using IQ-TREE 2 [66] with extended model selection followed by tree inference (-m MFP), 1000 ultrafast bootstrap replicates (-B 1000), 1000 replicates for single-branch testing (--alrt 1000), and ‘DNA’ specified as sequence type (--seqtype DNA). The phylogenetic tree was visualised and further modified using iTOL (https://itol.embl.de/, accessed on 14 June 2021) [67].

5.3. Isolation and Cloning of VviTPS04 and -10 Paralogues

Primers were designed to target putative paralogues: forward primer A (VviTPS10 paralogues): 5′-ATGGCCTTAATTCTCGCTACCAGCAACGGG-3′, forward primer B (VviTPS04 paralogues): 5′-ATGTTGAGTGCTCGAGTGAGTTTACACATGGCC-3′), and a single conserved reverse primer: 5′-TCATATTGGCACAGGGTCTATGAGCAGCATTGAAATATT-3′.
RNA extractions from 100-mg flower tissue of nine V. vinifera genotypes were performed using the Spectrum Plant Total RNA Kit (Sigma-Aldrich, St. Louis, MO, USA) with an on-column DNase I (Thermo Fisher Scientific, Waltham, MA, USA) treatment. RNA was checked for genomic DNA (gDNA) contamination through PCR, followed by reverse strand synthesis (cDNA) using the SuperScript VILO Master Mix (Thermo Fisher Scientific, USA). Genomic DNA (gDNA) was isolated from the same genotypes according to the method described by reference [68]. RNA extractions from 100-mg flower tissue of nine V. vinifera genotypes were performed using the Spectrum Plant Total RNA Kit (Sigma-Aldrich, St. Louis, MO, USA) with an on-column DNase I (Thermo Fisher Scientific, Waltham, MA, USA) treatment. RNA was checked for genomic DNA (gDNA) contamination through PCR, followed by reverse strand synthesis (cDNA) using the SuperScript VILO Master Mix (Thermo Fisher Scientific, Waltham, MA, USA). Genomic DNA (gDNA) was isolated from the same genotypes according to the method described by reference [68]. Flower genotypes were collected at stage 17/18 according to the modified Eichhorn and Lorenz (EL) growth classification system [69] for the following cultivars: Chardonnay (CH), Chenin Blanc (CB), Muscat of Alexandria (MA), Pinot noir (PN), Pinotage (PI), Sauvignon Blanc (SB), Shiraz (SH), Viognier (VG), and Weisser Riesling (WR). Samples were a composite of six to eight flowers clusters that were frozen in liquid nitrogen, with the rachis structures separated before homogenisation. All genotypes were planted in close geographical proximity at a mother block in the Stellenbosch area (33°57′33.50″ S, 18°51′38.09″ E), South Africa.
PCR reactions using cDNA or gDNA as the template was performed using TaKaRa ExTaq proofreading polymerase, as per the product specifications (Separations, Roodepoort, South Africa). PCR reaction conditions were as follows: 2 min initial denaturation, followed by 35 cycles of 30 sec denaturation, 30 s annealing, 2 min of 30-s extension, and a final extension of 7 min. cDNA amplicons were purified after electrophoretic separation using the Zymoclean Gel DNA Recovery Kit (Inqaba Biotech, Pretoria, South Africa), TA-cloned, screened for positives through colony PCR, and the plasmids isolated from overnight cultures. Plasmid DNA was diluted to 1/50 and used as a template for two-step Gateway PCR with the primers modified accordingly to build in attB sites, as described in the Gateway Technology Manual (Thermo Fisher Scientific, Waltham, MA, USA). Gateway cloning was performed according to the aforementioned manual using the appropriate clonases to generate entry clones with pDONR-Zeocin (Thermo Fisher Scientific, Waltham, MA, USA) and expression clones with pEAQ-HT-DEST1 [70,71]. The respective clones were transformed into chemically competent DH5α Escherichia coli using appropriate antibiotics and screened for positives through colony PCR. Entry clones were bidirectionally sequenced according to the standard methods of the Central Analytical Facility, Stellenbosch University, South Africa using M13 sequencing primers and a walking primer (5′-CTCTTCTATTGTTGGTATGTATTTTC-3′).

5.4. Sequence Analysis of Isolated Paralogues

Sequence analyses were performed using CLC Main Workbench 7 (CLC Bio-Qiagen, Aarhus, Denmark). Phased diploid annotations of the VviTPS family [27] were queried to identify the cluster of proteins most similar to the sequenced isolates. MUSCLE alignments [72,73] and phylogenetic tree construction were performed as described earlier using CLC Main Workbench 7 (CLC Bio-Qiagen, Aarhus, Denmark). Gene sequence data and Cytoscape network data were retrieved, as per the instructions in reference [27]. The I′ score [38] was calculated from BLASTn [62] scores of isolates queried against all diploid complete VviTPS genes (mRNA sequences) predicted by reference [27].

5.5. Gene Synteny Analysis

The MCScan pipeline [74] in the JCVI utility library [75] was used to perform a gene synteny analysis. Coding sequences for genes on contigs VvCabSauv08_v1_Primary000071F, VvCabSauv08_v1_Primary000049F [26], 000266F of Chardonnay [24], and PN40024 chr. 18 [1,21] were retrieved. The corresponding gene annotations were used to parse the GFF files to the BED format required for jcvi:MCScan. The gene synteny (microsynteny) analysis and visualisation were performed according to the instructions in the JCVI utility library using the default settings.

5.6. Agrobacterium-Mediated Transient Expression

The plasmids listed were transformed via electroporation into the Agrobacterium tumefaciens GV3101::pMP90 strain, as described by reference [76], and plated onto selective media. Agrobacterium-mediated transient expression in Nicotiana benthamiana was performed, as described by reference [20], with three leaves from a single plant pooled after four days for the subsequent analysis. A mock infiltration using MMA buffer and non-infiltrated wild-type leaves served as controls. Headspace solid-phase microextraction (HS-SPME) and gas chromatography-mass spectrometry (GC-MS) was performed as described by reference [77], with the following modifications: adsorption of analytes was done at 30 °C for 20 min and an initial oven temperature of 100 °C. Sesquiterpenes were identified using the authentic standards (E)-β-farnesene and (E)-β-caryophyllene from Sigma-Aldrich, USA and/or the Wiley 275 mass spectral library. For extracted ion chromatograms (EIC), the cumulative mass (m/z) response for 161, 189, and 204 was used.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/plants10081520/s1: Supplementary Figure S1: Complete phylogenetic tree of the gene models predicted from WGRS mappings with the primer bindings indicated. Supplementary File S1: Nucleotide sequences for reference genes and those predicted by mapping of the WGRS data. Supplementary File S2: Pairwise alignments of genotypic variants isolated from different V. vinifera genotypes. Supplementary File S3: Amino acid alignments of the phylogenetic tree shown in Figure 3B. Supplementary File S4: Amino acid alignments of the functional genotypic variants.

Author Contributions

Conceptualisation, S.J.S. and P.R.Y.; methodology, S.J.S.; software, S.J.S.; validation, S.J.S.; formal analysis, S.J.S.; investigation, S.J.S.; resources, S.J.S.; data curation, S.J.S.; writing—original draft preparation, S.J.S.; writing—review and editing, S.J.S., M.A.V., and P.R.Y.; visualisation, S.J.S.; supervision, M.A.V. and P.R.Y.; project administration, P.R.Y.; and funding acquisition, M.A.V. and P.R.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Wine Industry Network for Expertise and Technology (Winetech; IWBT P14/02), the National Research Foundation (NRF), Thuthuka (TTK13070220277), the Technology and Human Resources for Industry Programme (THRIP) of the Department of Trade and Industry (DTI) and a Consolidoc Fellowship from Stellenbosch University (2020).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented is contained within article and Supplementary Files.

Acknowledgments

The authors would like to recognise the following people for their contributions toward this study: Charles Visser from Vititec for allowing access to their vineyards to conduct field experiments. Franscisco Campos and Kalem Garcia for their assistance with sample collection and processing. Cody Williams for assistance with the implementation of the GC-MS methods.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of the data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Jaillon, O.; Aury, J.-M.; Noel, B.; Policriti, A.; Clepet, C.; Casagrande, A.; Choisne, N.; Aubourg, S.; Vitulo, N.; Jubin, C.; et al. The grapevine genome sequence suggests ancestral hexaploidization in major angiosperm phyla. Nature 2007, 449, 463–467. [Google Scholar] [CrossRef]
  2. Lin, J.; Massonnet, M.; Cantu, D. The genetic basis of grape and wine aroma. Hortic. Res. 2019, 6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Black, C.A.; Parker, M.; Siebert, T.E.; Capone, D.L.; Francis, I.L. Terpenoids and their role in wine flavour: Recent advances. Aust. J. Grape Wine Res. 2015, 21. [Google Scholar] [CrossRef]
  4. Skinkis, P.A.; Bordelon, B.P.; Wood, K.V. Comparison of monoterpene constituents in Traminette, Gewürztraminer, and Riesling winegrapes. Am. J. Enol. Vitic. 2008, 59, 440–445. [Google Scholar]
  5. Wood, C.; Siebert, T.E.; Parker, M.; Capone, D.L.; Elsey, G.M.; Pollnitz, A.P.; Eggers, M.; Meier, M.; Vössing, T.; Widder, S.; et al. From wine to pepper: Rotundone, an obscure sesquiterpene, is a potent spicy aroma compound. J. Agric. Food Chem. 2008, 56, 3738–3744. [Google Scholar] [CrossRef] [PubMed]
  6. Kalua, C.M.; Boss, P.K. Evolution of volatile compounds during the development of cabernet sauvignon grapes (Vitis vinifera L.). J. Agric. Food Chem. 2009, 57, 3818–3830. [Google Scholar] [CrossRef]
  7. Martin, D.; Aubourg, S.; Schouwey, M.; Daviet, L.; Schalk, M.; Toub, O.; Lund, S.; Bohlmann, J. Functional Annotation, Genome Organization and Phylogeny of the Grapevine (Vitis vinifera) Terpene Synthase Gene Family Based on Genome Assembly, FLcDNA Cloning, and Enzyme Assays. BMC Plant Biol. 2010, 10, 226. [Google Scholar] [CrossRef] [Green Version]
  8. Bloch, K.; Chaykin, S.; Phillips, A.; De Waard, A. Mevalonic acid pyrophospahte and isopentenylpyrphosphate. J. Biol. Chem. 1959, 234, 2595–2604. [Google Scholar] [CrossRef]
  9. Lichtenthaler, H.K. The 1-deoxy-D-xylulose-5-phosphate pathway of isoprenoid biosynthesis in plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 47–65. [Google Scholar] [CrossRef]
  10. Rohmer, M. A Mevalonate-independent Route to Isopentenyl Diphosphate. In Comprehensive Natural Product Chemistry. Isoprenoids Including Carotenoids and Steroids; Cane, D.E., Ed.; Elsevier: Amsterdam, The Netherlands, 1999; Volume 2, pp. 45–65. ISBN 0-08-043154-2. [Google Scholar]
  11. Davis, E.M.; Croteau, R. Cyclization enzymes in the biosynthesis of monoterpenes, sesquiterpenes, and diterpenes. Top. Curr. Chem. 2000, 209, 53–95. [Google Scholar]
  12. Martin, D.M.; Bohlmann, J. Identification of Vitis vinifera (-)-α-terpineol synthase by in silico screening of full-length cDNA ESTs and functional characterization of recombinant terpene synthase. Phytochemistry 2004, 65, 1223–1229. [Google Scholar] [CrossRef]
  13. Yoshikuni, Y.; Ferrin, T.E.; Keasling, J.D. Designed divergent evolution of enzyme function. Nature 2006, 440, 1078–1082. [Google Scholar] [CrossRef] [PubMed]
  14. Li, J.-X.; Fang, X.; Zhao, Q.; Ruan, J.-X.; Yang, C.-Q.; Wang, L.-J.; Miller, D.J.; Faraldos, J.A.; Allemann, R.K.; Chen, X.-Y.; et al. Rational engineering of plasticity residues of sesquiterpene synthases from Artemisia annua: Product specificity and catalytic efficiency. Biochem. J. 2013, 451, 417–426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Leferink, N.G.H.; Ranaghan, K.E.; Karuppiah, V.; Currin, A.; van der Kamp, M.W.; Mulholland, A.J.; Scrutton, N.S. Experiment and Simulation Reveal How Mutations in Functional Plasticity Regions Guide Plant Monoterpene Synthase Product Outcome. ACS Catal. 2018, 3780–3791. [Google Scholar] [CrossRef] [PubMed]
  16. Cane, D.E. Enzymic formation of sesquiterpenes. Chem. Rev. 1990, 90, 1089–1103. [Google Scholar] [CrossRef]
  17. Christianson, D.W. Structural biology and chemistry of the terpenoid cyclases. Chem. Rev. 2006, 106, 3412–3442. [Google Scholar] [CrossRef] [PubMed]
  18. Wedler, H.; Pemberton, R.; Tantillo, D. Carbocations and the complex flavor and bouquet of wine: Mechanistic aspects of terpene biosynthesis in wine grapes. Molecules 2015, 20, 10781–10792. [Google Scholar] [CrossRef] [Green Version]
  19. Drew, D.P.; Andersen, T.B.; Sweetman, C.; Møller, B.L.; Ford, C.; Simonsen, H.T. Two key polymorphisms in a newly discovered allele of the Vitis vinifera TPS24 gene are responsible for the production of the rotundone precursor α-guaiene. J. Exp. Bot. 2015, 67, 799–808. [Google Scholar] [CrossRef] [Green Version]
  20. Smit, S.J.; Vivier, M.A.; Young, P.R. Linking terpene synthases to sesquiterpene metabolism in grapevine flowers. Front. Plant Sci. 2019, 10, 1–18. [Google Scholar] [CrossRef] [Green Version]
  21. Canaguier, A.; Grimplet, J.; Gaspero, G.D.; Scalabrin, S.; Duchêne, E.; Choisne, N. A new version of the grapevine reference genome assembly (12X.v2) and of its annotation (VCost.v3). Genom. Data 2017, 14, 56–62. [Google Scholar] [CrossRef]
  22. Figueroa-Balderas, R.; Minio, A.; Morales-Cruz, A.; Vondras, A.M.; Cantu, D. Strategies for sequencing and assembling grapevine genomes. In The Grape Genome; Cantu, D., Walker, M.A., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 77–88. [Google Scholar]
  23. Minio, A.; Massonnet, M.; Figueroa-Balderas, R.; Vondras, A.M.; Blanco-Ulate, B.; Cantu, D. Iso-Seq Allows Genome-Independent Transcriptome Profiling of Grape Berry Development. G3 Genes Genomes Genet. 2019, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Roach, M.J.; Johnson, D.L.; Bohlmann, J.; van Vuuren, H.J.J.; Jones, S.J.M.; Pretorius, I.S.; Schmidt, S.A.; Borneman, A.R. Population sequencing reveals clonal diversity and ancestral inbreeding in the grapevine cultivar Chardonnay. PLoS Genet. 2018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Minio, A.; Massonnet, M.; Figueroa-Balderas, R.; Castro, A.; Cantu, D. Diploid Genome Assembly of the Wine Grape Carménère. G3 Genes Genomes Genet. 2019, 9. [Google Scholar] [CrossRef] [Green Version]
  26. Chin, C.; Peluso, P.; Sedlazeck, F.J.; Nattestad, M.; Concepcion, G.T.; Clum, A.; Dunn, C.; O’Malley, R.; Figueroa-Balderas, R.; Morales-Cruz, A.; et al. Phased diploid genome assembly with single-molecule real-time sequencing. Nat. Methods 2016, 13, 1050–1054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Smit, S.J.; Vivier, M.A.; Young, P.R. Comparative (within species) genomics of the Vitis vinifera L. terpene synthase family to explore the impact of genotypic variation using phased diploid genomes. Front. Genet. 2020, 11. [Google Scholar] [CrossRef]
  28. Carretero-Paulet, L.; Librado, P.; Chang, T.H.; Ibarra-Laclette, E.; Herrera-Estrella, L.; Rozas, J.; Albert, V.A. High gene family turnover rates and gene space adaptation in the compact genome of the carnivorous plant utricularia gibba. Mol. Biol. Evol. 2015, 32, 1284–1295. [Google Scholar] [CrossRef] [Green Version]
  29. Del Pozo, J.C.; Ramirez-Parra, E. Whole genome duplications in plants: An overview from Arabidopsis. J. Exp. Bot. 2015, 66, 6991–7003. [Google Scholar] [CrossRef] [Green Version]
  30. Giannuzzi, G.; D’Addabbo, P.; Gasparro, M.; Martinelli, M.; Carelli, F.N.; Antonacci, D.; Ventura, M. Analysis of high-identity segmental duplications in the grapevine genome. BMC Genom. 2011, 12, 436. [Google Scholar] [CrossRef] [Green Version]
  31. Aradhya, M.K.; Dangl, G.S.; Prins, B.H.; Boursiquot, J.-M.; Walker, M.A.; Meredith, C.P.; Simon, C.J. Genetic structure and differentiation in cultivated grape, Vitis vinifera L. Genet. Res. 2003, 81, 179–192. [Google Scholar] [CrossRef] [Green Version]
  32. Salmaso, M.; Faes, G.; Segala, C.; Stefanini, M.; Salakhutdinov, I.; Zyprian, E.; Toepfer, R.; Grando, M.S.; Velasco, R. Genome diversity and gene haplotypes in the grapevine (Vitis vinifera L.), as revealed by single nucleotide polymorphisms. Mol. Breed. 2004, 14, 385–395. [Google Scholar] [CrossRef]
  33. Laucou, V.; Launay, A.; Bacilieri, R.; Lacombe, T.; Adam-Blondon, A.F.; Bérard, A.; Chauveau, A.; De Andrés, M.T.; Hausmann, L.; Ibáñez, J.; et al. Extended diversity analysis of cultivated grapevine Vitis vinifera with 10K genome-wide SNPs. PLoS ONE 2018, 13, e192540. [Google Scholar] [CrossRef]
  34. Zhou, Y.; Muyle, A.; Gaut, B.S. Evolutionary genomics and the domestication of grapes. In The Grape Genome; Cantu, D., Walker, M.A., Eds.; Springer International Publishing: Cham, Swizterland, 2019; pp. 39–55. [Google Scholar]
  35. Falara, V.; Akhtar, T.A.; Nguyen, T.T.H.H.; Spyropoulou, E.A.; Bleeker, P.M.; Schauvinhold, I.; Matsuba, Y.; Bonini, M.E.; Schilmiller, A.L.; Last, R.L.; et al. The tomato terpene synthase gene family. Plant Physiol. 2011, 157, 770–789. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Chen, H.; Köllner, T.G.; Li, G.; Wei, G.; Chen, X.; Zeng, D.; Qian, Q.; Chen, F. Combinatorial evolution of a terpene synthase gene cluster explains terpene variations in oryza. Plant Physiol. 2019. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Liang, Z.; Duan, S.; Sheng, J.; Zhu, S.; Ni, X.; Shao, J.; Liu, C.; Nick, P.; Du, F.; Fan, P.; et al. Whole-genome resequencing of 472 Vitis accessions for grapevine diversity and demographic history. Nat. Commun. 2019, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Li, W.H.; Gu, Z.; Wang, H.; Nekrutenko, A. Evolutionary analyses of the human genome. Nature 2001, 409, 847–849. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Dueholm, B.; Drew, D.P.; Sweetman, C.; Simonsen, H.T. In planta and in silico characterization of five sesquiterpene synthases from Vitis vinifera (cv. Shiraz) berries. Planta 2019, 249, 59–70. [Google Scholar] [CrossRef] [Green Version]
  40. Myles, S.; Chia, J.M.; Hurwitz, B.; Simon, C.; Zhong, G.Y.; Buckler, E.; Ware, D. Rapid genomic characterization of the genus Vitis. PLoS ONE 2010, 5, e0008219. [Google Scholar] [CrossRef] [Green Version]
  41. Zhou, Y.; Massonnet, M.; Sanjak, J.S.; Cantu, D.; Gaut, B.S. Evolutionary genomics of grape (Vitis vinifera ssp. vinifera) domestication. Proc. Natl. Acad. Sci. USA 2017, 114, 11715–11720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Tabidze, V.; Pipia, I.; Gogniashvili, M.; Kunelauri, N.; Ujmajuridze, L.; Pirtskhalava, M.; Vishnepolsky, B.; Hernandez, A.G.; Fields, C.J.; Beridze, T. Whole genome comparative analysis of four Georgian grape cultivars. Mol. Genet. Genom. 2017, 292, 1377–1389. [Google Scholar] [CrossRef]
  43. Mercenaro, L.; Nieddu, G.; Porceddu, A.; Pezzotti, M.; Camiolo, S. Sequence polymorphisms and structural variations among four grapevine (Vitis vinifera L.) cultivars representing sardinian agriculture. Front. Plant Sci. 2017, 8, 1279. [Google Scholar] [CrossRef] [Green Version]
  44. Di Genova, A.; Almeida, A.M.; Muñoz-Espinoza, C.; Vizoso, P.; Travisany, D.; Moraga, C.; Pinto, M.; Hinrichsen, P.; Orellana, A.; Maass, A. Whole genome comparison between table and wine grapes reveals a comprehensive catalog of structural variants. BMC Plant Biol. 2014, 14. [Google Scholar] [CrossRef] [Green Version]
  45. Keller, M. Botany and anatomy. In The Science of Grapevines: Anatomy and Physiology; Academic Press: San Diego, CA, USA, 2010; pp. 1–47. ISBN 978-0-12-374881-2. [Google Scholar]
  46. Kronenberg, Z.N.; Rhie, A.; Koren, S.; Concepcion, G.T.; Peluso, P.; Munson, K.M.; Hiendleder, S.; Fedrigo, O.; Jarvis, E.D.; Adam, M.; et al. Extended haplotype phasing of de novo genome assemblies with FALCON-Phase. bioRxiv 2018. [Google Scholar] [CrossRef] [Green Version]
  47. Pichersky, E.; Raguso, R.A. Why do plants produce so many terpenoid compounds? New Phytol. 2018, 220, 692–702. [Google Scholar] [CrossRef]
  48. Myles, S.; Boyko, A.R.; Owens, C.L.; Brown, P.J.; Grassi, F.; Aradhya, M.K.; Prins, B.; Reynolds, A.; Chia, J.-M.; Ware, D.; et al. Genetic structure and domestication history of the grape. Proc. Natl. Acad. Sci. USA 2011, 108, 3530–3535. [Google Scholar] [CrossRef] [Green Version]
  49. Carbonell-Bejerano, P.; Royo, C.; Torres-Pérez, R.; Grimplet, J.; Fernandez, L.; Franco-Zorrilla, J.M.; Lijavetzky, D.; Baroja, E.; Martínez, J.; García-Escudero, E.; et al. Catastrophic unbalanced genome rearrangements cause somatic loss of berry color in grapevine. Plant Physiol. 2017, 175, 786–801. [Google Scholar] [CrossRef] [Green Version]
  50. Xu, Y.; Gao, Z.; Tao, J.; Jiang, W.; Zhang, S.; Wang, Q.; Qu, S. Genome-wide detection of SNP and SV variations to reveal early ripening-related genes in grape. PLoS ONE 2016, 11, e0147749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Cardone, M.F.; D’Addabbo, P.; Alkan, C.; Bergamini, C.; Catacchio, C.R.; Anaclerio, F.; Chiatante, G.; Marra, A.; Giannuzzi, G.; Perniola, R.; et al. Inter-varietal structural variation in grapevine genomes. Plant J. 2016, 88. [Google Scholar] [CrossRef] [PubMed]
  52. Zhou, Y.S.; Minio, A.; Massonnet, M.; Solares, E.A.; Lyu, Y.; Beridze, T.; Cantu, D.; Gaut, B.S. Structural variants, clonal propagation, and genome evolution in grapevine (Vitis vinifera). bioRxiv 2018. [Google Scholar] [CrossRef]
  53. Purugganan, M.D. Evolutionary insights into the nature of plant domestication. Curr. Biol. 2019, 29, R705–R714. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Olmo, H. Selecting and breeding new grape varieties. Calif. Agric. 1980, 34, 23–24. [Google Scholar]
  55. Pertot, I.; Caffi, T.; Rossi, V.; Mugnai, L.; Hoffmann, C.; Grando, M.S.; Gary, C.; Lafond, D.; Duso, C.; Thiery, D.; et al. A critical review of plant protection tools for reducing pesticide use on grapevine and new perspectives for the implementation of IPM in viticulture. Crop Prot. 2017, 97, 70–84. [Google Scholar] [CrossRef]
  56. Tasin, M.; Bäckman, A.C.; Bengtsson, M.; Ioriatti, C.; Witzgall, P. Essential host plant cues in the grapevine moth. Naturwissenschaften 2006, 93, 141–144. [Google Scholar] [CrossRef] [PubMed]
  57. Tasin, M.; Betta, E.; Carlin, S.; Gasperi, F.; Mattivi, F.; Pertot, I. Volatiles that encode host-plant quality in the grapevine moth. Phytochemistry 2011, 72, 1999–2005. [Google Scholar] [CrossRef]
  58. Salvagnin, U.; Malnoy, M.; Thöming, G.; Tasin, M.; Carlin, S.; Martens, S.; Vrhovsek, U.; Angeli, S.; Anfora, G. Adjusting the scent ratio: Using genetically modified Vitis vinifera plants to manipulate European grapevine moth behaviour. Plant Biotechnol. J. 2018, 16, 264–271. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Addison, P. Ecology and Management of Lobesia vanillana in Vineyards. Available online: https://www.wineland.co.za/new-winetech-projects-2017-part-2/ (accessed on 17 December 2019).
  60. Picaud, S.; Brodelius, M.; Brodelius, P.E. Expression, purification and characterization of recombinant (E)-β-farnesene synthase from Artemisia annua. Phytochemistry 2005, 66, 961–967. [Google Scholar] [CrossRef]
  61. Dèrozier, S.; Samson, F.; Tamby, J.-P.; Guichard, C.; Brunaud, V.; Grevet, P.; Gagnot, S.; Label, P.; Leplé, J.-C.; Lecharny, A.; et al. Exploration of plant genomes in the FLAGdb++ environment. Plant Methods 2011, 7, 8. [Google Scholar] [CrossRef] [Green Version]
  62. Camacho, C.; Coulouris, G.; Avagyan, V.; Ma, N.; Papadopoulos, J.; Bealer, K.; Madden, T.L. BLAST+: Architecture and applications. BMC Bioinform. 2009, 10, 421. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Boratyn, G.M.; Thierry-Mieg, J.; Thierry-Mieg, D.; Busby, B.; Madden, T.L. Magic-BLAST, an accurate RNA-seq aligner for long and short reads. BMC Bioinform. 2019, 20. [Google Scholar] [CrossRef]
  64. Sievers, F.; Wilm, A.; Dineen, D.; Gibson, T.J.; Karplus, K.; Li, W.; Lopez, R.; McWilliam, H.; Remmert, M.; Söding, J.; et al. Fast, scalable generation of high-quality protein multiple sequence alignments using Clustal Omega. Mol. Syst. Biol. 2011, 7. [Google Scholar] [CrossRef]
  65. Sievers, F.; Higgins, D.G. Clustal Omega for making accurate alignments of many protein sequences. Protein Sci. 2018, 27, 135–145. [Google Scholar] [CrossRef] [Green Version]
  66. Minh, B.Q.; Schmidt, H.A.; Chernomor, O.; Schrempf, D.; Woodhams, M.D.; Von Haeseler, A.; Lanfear, R.; Teeling, E. IQ-TREE 2: New models and efficient methods for phylogenetic inference in the genomic era. Mol. Biol. Evol. 2020, 37, 1530–1534. [Google Scholar] [CrossRef] [Green Version]
  67. Letunic, I.; Bork, P. Interactive Tree Of Life (iTOL) v5: An online tool for phylogenetic tree display and annotation. Nucleic Acids Res. 2021, 49, W293–W296. [Google Scholar] [CrossRef]
  68. Reid, K.E.; Olsson, N.; Schlosser, J.; Peng, F.; Lund, S.T. An optimized grapevine RNA isolation procedure and statistical determination of reference genes for real-time RT-PCR during berry development. BMC Plant Biol. 2006, 6, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Coombe, B.G. Adoption of a system for identifying grapevine growth stages. Aust. J. Grape Wine Res. 1995, 1, 100–110. [Google Scholar] [CrossRef]
  70. Peyret, H.; Lomonossoff, G.P. The pEAQ vector series: The easy and quick way to produce recombinant proteins in plants. Plant Mol. Biol. 2013, 83, 51–58. [Google Scholar] [CrossRef] [PubMed]
  71. Sainsbury, F.; Thuenemann, E.C.; Lomonossoff, G.P. pEAQ: Versatile expression vectors for easy and quick transient expression of heterologous proteins in plants. Plant Biotechnol. J. 2009, 7, 682–693. [Google Scholar] [CrossRef] [PubMed]
  72. Edgar, R.C. MUSCLE: A multiple sequence alignment method with reduced time and space complexity. BMC Bioinform. 2004, 5. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Edgar, R.C. MUSCLE: Multiple sequence alignment with high accuracy and high throughput. Nucleic Acid Res. 2004, 32, 1792–1797. [Google Scholar] [CrossRef] [Green Version]
  74. Tang, H.; Bowers, J.E.; Wang, X.; Ming, R.; Alam, M.; Paterson, A.H. Synteny and collinearity in plant genomes. Science 2008, 320, 486–488. [Google Scholar] [CrossRef] [Green Version]
  75. Tang, H.; Krishnakumar, V.; Li, J. jcvi: JCVI utility libraries. Zenodo 2015. [Google Scholar] [CrossRef]
  76. Bach, S.S.; Bassard, J.-É.; Andersen-Ranberg, J.; Møldrup, M.E.; Simonsen, H.T.; Hamberger, B. High-throughput testing of terpenoid biosynthesis candidate genes using transient expression in Nicotiana benthamiana. In Plant Isoprenoids: Methods and Protocols; Rodríguez-Concepción, M., Ed.; Springer Science: New York, NY, USA, 2014; Volume 1153, pp. 245–255. ISBN 978-1-4939-0605-5. [Google Scholar]
  77. Williams, C.; Buica, A. Comparison of an offline-SPE-GC-MS and online-HS-SPME-GC-MS method for the analysis of volatile terpenoids in wine. Molecules 2019, 25, 657. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Phylogenetic tree of the consensus sequences assembled from the WGRS data [37] using VviTPS gDNA models predicted from phased diploid genomes [27] for the cultivars Chardonnay (CH) [24], Carménère (CR) [25], and Cabernet Sauvignon (CS) [26] and the grapevine reference genome PN40024 [1,7]. The PN40024 VviTPS04 and -10 gDNA models are indicated in bold. WGRS genotypes for which the gene models were predicted are shown, followed by the gDNA model [27] that the reads mapped to. The coloured circles show which primers bind at the 5′ terminal ends. Triangular symbols on the tree indicate high confidence branch nodes with an approximate likelihood-ratio test (SH-aLRT) >85% and ultrafast bootstrap support >95% using 1000 replicates.
Figure 1. Phylogenetic tree of the consensus sequences assembled from the WGRS data [37] using VviTPS gDNA models predicted from phased diploid genomes [27] for the cultivars Chardonnay (CH) [24], Carménère (CR) [25], and Cabernet Sauvignon (CS) [26] and the grapevine reference genome PN40024 [1,7]. The PN40024 VviTPS04 and -10 gDNA models are indicated in bold. WGRS genotypes for which the gene models were predicted are shown, followed by the gDNA model [27] that the reads mapped to. The coloured circles show which primers bind at the 5′ terminal ends. Triangular symbols on the tree indicate high confidence branch nodes with an approximate likelihood-ratio test (SH-aLRT) >85% and ultrafast bootstrap support >95% using 1000 replicates.
Plants 10 01520 g001
Figure 2. The gDNA PCR of nine V. vinifera genotypes using the conserved reverse primer in combination with primer A (A) or primer B (B). The following abbreviations were used for the respective genotypes: Cabernet Sauvignon (CS), Chardonnay (CH), Chenin Blanc (CB), Muscat of Alexandria (MA), Pinot noir (PN), Sauvignon Blanc (SB), Shiraz (SH), and Viognier (VI). The 27-bp difference in the VviTPS04 and VviTPS10 reference genome gene models with the binding positions of their respective forward primers are shown (C). The two in-frame start codons are indicated on VviTPS04 by the red arrows.
Figure 2. The gDNA PCR of nine V. vinifera genotypes using the conserved reverse primer in combination with primer A (A) or primer B (B). The following abbreviations were used for the respective genotypes: Cabernet Sauvignon (CS), Chardonnay (CH), Chenin Blanc (CB), Muscat of Alexandria (MA), Pinot noir (PN), Sauvignon Blanc (SB), Shiraz (SH), and Viognier (VI). The 27-bp difference in the VviTPS04 and VviTPS10 reference genome gene models with the binding positions of their respective forward primers are shown (C). The two in-frame start codons are indicated on VviTPS04 by the red arrows.
Plants 10 01520 g002
Figure 3. cDNA sequences isolated from different genotypes using the two different primer combinations shown (A). Phylogenetic positions relative to the VvGWaBer and VviMATPS10 nucleotide sequences are shown (B). The letter code in parentheses refers to the primer set used for isolation with the type of disruption indicated when non-functional.
Figure 3. cDNA sequences isolated from different genotypes using the two different primer combinations shown (A). Phylogenetic positions relative to the VvGWaBer and VviMATPS10 nucleotide sequences are shown (B). The letter code in parentheses refers to the primer set used for isolation with the type of disruption indicated when non-functional.
Plants 10 01520 g003
Figure 4. (A) Cumulative EIC using the masses 161, 189, and 204 for N. benthamiana transiently expressing genes of interest. Cultivar-specific genes are shown with peaks identified using authentic standards (STD). Wild-type (WT) and mock infiltrations served as controls. The m/z spectra for the peaks indicated by the asterisk are shown (B), with a dashed grey line showing that its retention time is different to that of (E)-β-farnesene. The following cultivar abbreviations were used: Chenin Blanc (CB), Cabernet Sauvignon (CS), Muscat of Alexandria (MA), Pinot noir (PN), and Viognier (VI).
Figure 4. (A) Cumulative EIC using the masses 161, 189, and 204 for N. benthamiana transiently expressing genes of interest. Cultivar-specific genes are shown with peaks identified using authentic standards (STD). Wild-type (WT) and mock infiltrations served as controls. The m/z spectra for the peaks indicated by the asterisk are shown (B), with a dashed grey line showing that its retention time is different to that of (E)-β-farnesene. The following cultivar abbreviations were used: Chenin Blanc (CB), Cabernet Sauvignon (CS), Muscat of Alexandria (MA), Pinot noir (PN), and Viognier (VI).
Plants 10 01520 g004
Figure 5. VviTPS network connectivity [27] for the paralogues targeted in this study. Genotype-specific duplication networks are shown (A). The best-scoring complete gene match of the isolated genes is shown (B), with the edge values representing the I′ score. (C) The protein cluster associated with the isolates. The following cultivar abbreviations were used: Chenin Blanc (CB), Chardonnay (CH), Carménère (CR), Cabernet Sauvignon (CS), Muscat of Alexandria (MA), Pinot noir (PN), and Viognier (VI).
Figure 5. VviTPS network connectivity [27] for the paralogues targeted in this study. Genotype-specific duplication networks are shown (A). The best-scoring complete gene match of the isolated genes is shown (B), with the edge values representing the I′ score. (C) The protein cluster associated with the isolates. The following cultivar abbreviations were used: Chenin Blanc (CB), Chardonnay (CH), Carménère (CR), Cabernet Sauvignon (CS), Muscat of Alexandria (MA), Pinot noir (PN), and Viognier (VI).
Plants 10 01520 g005
Figure 6. Gene synteny of the VviTPS landscape for the Cabernet Sauvignon (CS) and Chardonnay (CH) contigs relative to the PN40024 chr. 18 VviTPS cluster. The arrows on the contigs/chr.18 show the direction of the genes, with the VviTPS genes shown in light blue. The genes on the CS and CH contigs that are orthologous to PN40024 VviTPS04 and VviTPS10 are connected by purple and gold syntenic lines, respectively.
Figure 6. Gene synteny of the VviTPS landscape for the Cabernet Sauvignon (CS) and Chardonnay (CH) contigs relative to the PN40024 chr. 18 VviTPS cluster. The arrows on the contigs/chr.18 show the direction of the genes, with the VviTPS genes shown in light blue. The genes on the CS and CH contigs that are orthologous to PN40024 VviTPS04 and VviTPS10 are connected by purple and gold syntenic lines, respectively.
Plants 10 01520 g006
Figure 7. Active site alignment of the six functional isolates to VviMATPS10 and VvGwaBer are shown in (A). The position of the DDxxD and NSE/DTE active site motifs are highlighted in green. The active site similarity is shown in the phylo-genetic tree (B). The following cultivar abbreviations were used: Chenin Blanc (CB), Cabernet Sauvignon (CS), Muscat of Alexandria (MA), Pinot noir (PN), and Viognier (VI).
Figure 7. Active site alignment of the six functional isolates to VviMATPS10 and VvGwaBer are shown in (A). The position of the DDxxD and NSE/DTE active site motifs are highlighted in green. The active site similarity is shown in the phylo-genetic tree (B). The following cultivar abbreviations were used: Chenin Blanc (CB), Cabernet Sauvignon (CS), Muscat of Alexandria (MA), Pinot noir (PN), and Viognier (VI).
Plants 10 01520 g007
Table 1. Consensus sequences assembled from WGRS data [37] using VviTPS gDNA models predicted from phased diploid genomes [27] for the cultivars Chardonnay (CH) [24], Carménère (CR) [25], and Cabernet Sauvignon (CS) [26] and the grapevine reference genome PN40024 [1,7]. Cultivars used in the mapping are indicated with their respective accessions, indicated in parentheses. Consensus sequences with a binding site for either primer A (green) or primer B (red) are shown, with group C (orange) indicating sequences that do not have coverage at the 5′-terminal ends.
Table 1. Consensus sequences assembled from WGRS data [37] using VviTPS gDNA models predicted from phased diploid genomes [27] for the cultivars Chardonnay (CH) [24], Carménère (CR) [25], and Cabernet Sauvignon (CS) [26] and the grapevine reference genome PN40024 [1,7]. Cultivars used in the mapping are indicated with their respective accessions, indicated in parentheses. Consensus sequences with a binding site for either primer A (green) or primer B (red) are shown, with group C (orange) indicating sequences that do not have coverage at the 5′-terminal ends.
gDNA ModelGenomeCabernet Franc (TA-183)Carménère (TA-254)Cabernet Sauvignon (TA-280)Chardonnay (TA-238)Merlot (TA-241)Muscat of Alexandria (TA-170)Pinot Noir (TA-379)Riesling (TA-329)Sangiovese (TA-334)Sauvignon Blanc (TA-417)Syrah (TA-242)
CHTPS045CH A A
CHTPS046CHAAAACAAAAAA
CHTPS062CH BBB
CHTPS065CH C AAA A
CHTPS067CH B CCC B
CHTPS147CHAAAAA AAAAA
CHTPS159CH B
CHTPS173CHCCCCCCCCCCC
CRTPS019CRAAA A AAAA
CRTPS081CRAAC A CAAA
CRTPS102CR C
CRTPS105CR
CRTPS106CRAAA A AC AC
CRTPS151CRAAAAAA AACA
CRTPS173CR A
CRTPS199CR AAA A
CRTPS200CR B CC B
CSTPS036CS AA A AA
CSTPS040CSAAAAAA AACA
CSTPS043CS A
CSTPS047CS CB CC
CSTPS051CS BBB
CSTPS059CS AA C
CSTPS079CS AA CA A
CSTPS080CS AA AAAAAA
CSTPS083CS C
CSTPS086CS B B
CSTPS090CS C A A
VviTPS04PN40024 BBB B
VviTPS10PN40025 C ACA AC
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Smit, S.J.; Vivier, M.A.; Young, P.R. Seeing the Forest through the (Phylogenetic) Trees: Functional Characterisation of Grapevine Terpene Synthase (VviTPS) Paralogues and Orthologues. Plants 2021, 10, 1520. https://doi.org/10.3390/plants10081520

AMA Style

Smit SJ, Vivier MA, Young PR. Seeing the Forest through the (Phylogenetic) Trees: Functional Characterisation of Grapevine Terpene Synthase (VviTPS) Paralogues and Orthologues. Plants. 2021; 10(8):1520. https://doi.org/10.3390/plants10081520

Chicago/Turabian Style

Smit, Samuel J., Melané A. Vivier, and Philip R. Young. 2021. "Seeing the Forest through the (Phylogenetic) Trees: Functional Characterisation of Grapevine Terpene Synthase (VviTPS) Paralogues and Orthologues" Plants 10, no. 8: 1520. https://doi.org/10.3390/plants10081520

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