Skip to main content

ORIGINAL RESEARCH article

Front. Plant Sci., 18 January 2023
Sec. Plant Breeding

High-density genetic linkage mapping reveals low stability of QTLs across environments for economic traits in Eucalyptus

  • 1Key Laboratory of National Forestry and Grassland Administration on Tropical Forestry Research, Research Institute of Tropical Forestry, Chinese Academy of Forestry, Guangzhou,  China
  • 2Commonwealth Scientific and Industrial Research Organisation (CRISO) Australian Tree Seed Centre, Canberra, ACT,  Australia

Introduction: Eucalyptus urophylla, E. tereticornis and their hybrids are the most important commercial forest tree species in South China where they are grown for pulpwood and solid wood production. Construction of a fine-scale genetic linkage map and detecting quantitative trait loci (QTL) for economically important traits linked to these end-uses will facilitate identification of the main candidate genes and elucidate the regulatory mechanisms.

Method: A high-density consensus map (a total of 2754 SNPs with 1359.18 cM) was constructed using genotyping by sequencing (GBS) on clonal progenies of E. urophylla × tereticornis hybrids. QTL mapping of growth and wood property traits were conducted in three common garden experiments, resulting in a total of 108 QTLs. A total of 1052 candidate genes were screened by the efficient combination of QTL mapping and transcriptome analysis.

Results: Only ten QTLs were found to be stable across two environments, and only one (qSG10Stable mapped on chromosome 10, and associated with lignin syringyl-to-guaiacyl ratio) was stable across all three environments. Compared to other QTLs, qSG10Stable explained a very high level of phenotypic variation (18.4–23.6%), perhaps suggesting that QTLs with strong effects may be more stably inherited across multiple environments. Screened candidate genes were associated with some transcription factor families, such as TALE, which play an important role in the secondary growth of plant cell walls and the regulation of wood formation.

Discussion: While QTLs such as qSG10Stable, found to be stable across three sites, appear to be comparatively uncommon, their identification is likely to be a key to practical QTL-based breeding. Further research involving clonally-replicated populations, deployed across multiple target planting sites, will be required to further elucidate QTL-by-environment interactions.

1 Introduction

Tree growth and wood properties are important traits in breeding programs for fast-growing forest trees including Eucalyptus. Quantitative trait locus (QTL) mapping has revealed important genetic regions that have influence on growth and wood property traits of major forest tree species (Hall et al., 2016; Mori et al., 2019). The interaction between growing environment and QTL expression and mapping in forest trees is complex, and gaining a better understanding of the phenomenon may have significant impacts on a forest tree breeding program (Rae et al., 2007). In crops planted in the field, mounting evidence suggests that the majority of QTLs do not regulate target traits across multiple environments (Shang et al., 2015; Raihan et al., 2016; He et al., 2021). Similarly, in forest trees, some studies have attempted to identify the stability of QTLs across multiple environments. For example, Westbrook et al. (2015) investigated QTLs regulating the stem resin channels by integrating genetic analyses across environments, ages, and populations of Pinus taeda. Mori et al. (2019) performed QTL mapping for growth and wood properties of Cryptomeria japonica across three environments. These studies both suggest that for long-lived forest species, QTLs for growth and wood property traits tend to be sensitive to environments.

Globally, Eucalyptus is one of the most important cultivated forest tree genera, providing fuelwood, fibre and timber (Ladiges et al., 2003). Due to its economic importance, QTL mapping for growth traits and wood properties has been intensively carried out (Gion et al., 2011; Bartholomé et al., 2013; Freeman et al., 2013). Some significant challenges to making practical use of QTLs have been identified (Grattapaglia and Kirst, 2008). Firstly, QTLs identified in study populations have not proven to be stable in eucalypt breeding populations. This is a particular issue for highly genetically diverse breeding populations that are managed by open pollination or large numbers of controlled-pollinated mating among members. The issue is less likely to be problematic where progeny of small sets of controlled pollinated crosses are made, for example to produce hybrid progeny from elite pure-species parents. This latter scenario is more likely in tropical regions, including major plantation growing regions such as Brazil and southern China, where hybrids and clonal selections are typically deployed (Grattapaglia and Kirst, 2008; Xie et al., 2017).

A second issue to making practical use of QTLs is that growth and wood property traits are quantitative (i.e., numerous genes of small effect are likely to contribute) and are subject to genotype-by-environment interactions (G × E) (Malan and Verryn, 1996; Brawner et al., 2013; Bush et al., 2015; de Araujo et al., 2019). Quantitative genetic studies have shown that G × E is more marked for growth traits than it is for wood properties (Downes et al., 1997), although both are significant. It is therefore not surprising that evidence for QTL × environment interactions has been found (Freeman et al., 2013; Westbrook et al., 2015; Mori et al., 2019), the problem being that QTLs identified from a study of a single site may not be identified at a second or subsequent sites. The challenge is therefore to find QTLs that are stable over multiple sites. Among these studies, Freeman et al. (2013) highlighted QTL stability in regulating growth and wood properties across different pedigrees and multi-environmental trials in E. globulus using control-pollinated families, however, the impact of environmental effects is likely to be harder to detect using non-clonal plants.

Despite the more-sophisticated analytical methods which are now available, classical genetic linkage maps remain important tools for QTL analysis and efficient programs for breeding (Li et al., 2020b). In recent decades, molecular markers such as SSR, DArT and SNP have been used to construct genetic maps for forest trees (Kullan et al., 2012; Pavy et al., 2017). Of these, SNP markers based on high-throughput techniques are gaining popularity due to the dramatically increased marker density that can be achieved per unit cost. Moreover, the cost of SNP genotyping continues to fall, with the development of high-throughput technology such as genotyping by sequencing (GBS) (Elshire et al., 2011). Recently, GBS technology has been widely applied to the genotyping of perennial woody plants, such as Eucalyptus (Brenton et al., 2019; Klápště et al., 2021), Populus (Carletti et al., 2016; Xia et al., 2018), Cinnamomum (Gong et al., 2021) and Pinus (Hirao et al., 2022), accelerating the molecular breeding progress in these species. The publication of the E. grandis reference genome (Myburg et al., 2014) has also significantly aided QTL mapping efforts, with high synteny between E. grandis and other common commercial eucalypts. In Eucalyptus, most of the previous QTL studies found few substantive associations between target traits and candidate genes due to the low resolution of genetic map, large QTL interval, or lack of reference genomic information (Freeman et al., 2009; Thumma et al., 2010; Freeman et al., 2013; Sumathi et al., 2018).

Due to advances in high-throughput sequencing technologies, researchers have conducted some transcriptome sequencing (RNA-seq) to discover functional genes and try to reveal molecular mechanisms of complex traits in Eucalyptus (Nakahama et al., 2018; Vaillant et al., 2018; Zhang et al., 2022). For example, Zhang et al. (2022) performed RNA-seq to identify the co-expressed genes during the callus maturation and shoot regeneration of Eucalyptus, and found a number of potentially functional genes associated with reproductive capacity. Studies in other woody plants, such as in Picea glauca (Laoué et al., 2021) and Eriobotrya japonica (Peng et al., 2022), have shown that the combination of QTL mapping and transcriptome data will significantly improve the efficiency of mining for functional genes and elucidating the genetic mechanism.

Although considerable progress has been made in dissecting the genetic mechanism of growth and wood property traits, much remains unknown in Eucalyptus. Economic trait QTLs, their stability and transcriptome activity across environments is an important area for future research. In this study, large-scale population GBS sequencing and small-scale RNA-seq on clonal progenies of E. urophylla × tereticornis hybrids were performed. We aimed to (1) construct high-density and high-quality genetic linkage maps, (2) improve the mapping accuracy of candidate genes for economic traits by combining QTL mapping and RNA-seq, and (3) investigate the stability of QTLs across different environments and explore the impact of environmental effect on the QTL mapping progenies.

2 Materials and methods

2.1 Plant material and phenotyping

The mapping population was developed from a control-pollinated cross between two excellent genotypes, E. urophylla (female, genotype UX-30) and E. tereticornis (male, genotype T43-05). Both were selected for rapid growth and pulp fibre property traits including lignin, cellulose, and hemi-cellulose. A total of 320 E. urophylla × tereticornis F1 hybrid clonal progenies resulting from the cross was described in detail in a previous report (Yang et al., 2018a). Replicated field plantings (at least three ramets per clone per site) including all 320 clonal, full-sib progenies at Gonghe (GH, 112°51’ E, 22°34’ N), and subsets of 180 and 164 progenies at Jijia (JJ, 101°01’ E, 20°21’ N) in Guangdong Province, and Yanxi (YX, 117°52’ E, 24°46’ N) in Fujian Province, China, respectively, were established in April 2006 (Figure 1A). Field trials at all three sites were designed as randomized complete blocks with single-tree plots spaced 2 m × 3 m. The number of block-replicates at Gonghe, Jijia, and Yanxi was four, four, and three, respectively. These trials did not contain the parental genotypes. In January 2014, tree height (H, m) and diameter at breast height (D, cm) were measured, and wood properties including basic density (BD, g cm−3), cellulose content (CC, %), hemi-cellulose content (HC, %), lignin content (LC, %), and lignin syringyl-to-guaiacyl ratio (SG) were assessed using near-infrared analysis (Yang et al., 2018a). To assess the selection of trees for RNA sampling, which was carried out on the basis of phenotypic values at Gonghe, data for each trait, pooled across the three sites, was analyzed using a linear mixed individual-tree model of the following form (following Baltunis et al., 2007):

FIGURE 1
www.frontiersin.org

Figure 1 Phenotypic data summary. (A) The distribution of clonal progenies among three sites. Boxplots (phenotypic data) of H (B), D (C), BD (D), CC (E), HC (F), LC (G), SG (H) of E. urophylla × tereticornis hybrids from Gonghe, Jijia and Yanxi. (I) Pearson’ correlation (r) coefficients between across-sites genotype best linear unbiased predictions (BLUPs) for pairs of traits and significance of two-sided test (***, ** and * are P < 0.001, P < 0.01 and P<0.05). BLUPs for each genotype, shown as bar plots in Figure S1, are given as deviations from the grand mean for each trait (Table S3).

yi=Xisi+Zigi+Zibi+Ziri+eiEquation 1

where for each trait, yi is a vector of phenotypic measurements on each trait indexed (i) by trial for single trait analyses across sites or by trait for bivariate analyses; bi is a vector of fixed effects; si and gi are vectors of random-effect estimates for genetic and within-site spatial effects, respectively; and ei is a vector of random residual effects. Xi is the incidence matrix relating the observations in yi to the fixed effects for mean and site in bi:

Xisi=[X100X2][s1s2]Equation 2

gi is the vector of random genetic effects of individual genotypes ~MVN(0, G⊗A) where

G=[σg12σg1g2σg1g2σg22]Equation 3

and A is the pedigree-based numerator relationship matrix, Zi is the incidence matrix relating the observations in yi to the random genetic effects in gi, σ^g12 is the genetic variance and σg1g2 is the genetic covariance across sites. A comprised elements equivalent to twice the coefficient of kinship (Ɵ); with 2θ = 0.5 on the off-diagonals and 1+f on the diagonal (where f is each individual’s inbreeding coefficient, assumed to be zero), reflecting that each clonal progeny is resultant of a controlled pollination between two parents. Vector bi comprised random effects of complete block replicates within sites MVN(0,Ibiσbi2) where Ibi is an identity matrix corresponding to the number of blocks and σbi2 is the variance associated with the block replicates and analogously, ri is a random term associated with the planting rows within sites MVN(0,Ibiσbi2) where Iri is an identity matrix corresponding to the number of rows and σbi2  is the variance associated with the rows. Term ei is the random vector of residuals.

∼MVN(0[I1σe1200I2σe22])Equation 4

Where σe12  is the residual variance of each trait, I1 is the identity matrix corresponding to the number of trees at each site in the case of analysis of a single trait and 0 is the null matrix. For bivariate (trait-trait) correlations, correlated residuals were assumed.

The significance of the genotype effect was gauged by performing additional model runs omitting the term and comparing the Akaike Information Criterion between runs. The mixed models were solved using restricted maximum likelihood implemented in Asreml 4.2 (VSN International, Hemel Hempstead, UK). Bivariate genetic correlation estimates (rG) between traits x and y were obtained from the estimated genetic covariance and variance components as:

rG=σ^gxgyσgx2σgy2Equation 5

where σgxgy is the genetic covariance component between traits, and σgx2 and σgy2 are the genetic variance components for traits x and y respectively. These correlations were modelled using Asreml with simultaneous estimation of covariance, and separate genetic and error variances for each trait (see Butler et al., 2018). As solving the bivariate genetic correlation models was problematic for many combinations of traits, Pearson correlations between best linear unbiased predictions (BLUP) for pairs of traits on each genotype were additionally carried out using the FCORRELATION Procedure in Genstat 22 (VSN International, Hemel Hempstead, UK).

2.2 GBS library construction, sequencing, and SNP genotyping

Healthy leaves from 320 progenies in Gonghe population and the parental genotypes were collected. An improved CTAB method (Gan et al., 2003) was used for DNA extraction. The GBS library was constructed following the protocol of Poland et al. (2012), with some modifications. This involved screening and optimization of novel enzyme digestion combinations (MspI-MseI). The concentration of GBS library was quantified using a QUBIT fluorometer 3.0 (ThermoFisher, USA), and the size was assessed using the Labchip GXII Touch 24 (Perkin Elmer, USA). The qualified libraries were subjected to paired-end sequencing (PE 150) using the HiSeq 2500 platform (Illumina, USA). Low-quality reads (Q20<20, length<36bp) were subsequently filtered using Trimmomatic v0.36 software (Bolger et al., 2014). Reads were demultiplexed using the “process_radtags” programs of Stacks v2.49 (Rochette et al., 2019). After demultiplexing the raw data, construction of the Stacks catalogue, SNP calling, and genotype construction were performed for reference pipelines using Stacks ref_map.pl program aligned to the E. grandis reference genome v2.0 (https://phytozome.jgi.doe.gov/info/Egrandis_v2_0). SNP variants were annotated with SnpEff v4.3 (Cingolani et al., 2012). Finally, segregating SNP markers for a biparental control-pollinated mapping population with phase information from the parental genotypes were exported in the JoinMap v4.1 (Van Ooijen, 2006) format using the “genotypes” program. As a testcross between two heterozygous diploid parents that have resulted from the eucalypt mixed mating system, SNPs were defined as belonging to one of three segregation types: loci that segregate in a 1:1 ratio including those heterozygous in the male parent (E. tereticornis) and homozygous in the female (E. urophylla) and those heterozygous in the female parent and homozygous in the male); and those heterozygous in both parents segregating in a 1:2:1 ratio (see Maliepaard et al., 1997; Wu et al., 2002 for additional theory).

2.3 Genetic linkage map construction

The linkage map was constructed using JoinMap v4.1 and Lep-Map3 (Rastas et al., 2013). Firstly, JoinMap v4.1 was used to divide markers into groups using a grouping tree function in cross-pollinated model with LOD = 12.0 (Van Ooijen, 2006). Subsequently, Lep-Map3 was used to construct the parental genetic maps (Rastas et al., 2013). The process was as follows: Step 1: ParentCall2 was used to call parental genotypes, followed by the Filtering2 step which filtered markers based on high segregation distortion. The default value of the data tolerance (P=0.01) was used to discard highly segregated markers (χ2 test, P<0.01). Step 2: the SeparateChromosomes2 module was used to assigned markers into linkage groups (LGs) by computing all pairwise LOD scores between markers and joined markers with LOD scores range from four to ten. Step 3: OrderMarkers2 module was used to order the markers within each LG by maximizing the likelihood of the data for alternative orders. The Kosambi mapping function was used to estimate map distances. The consensus map for both the maternal and paternal maps was computed using the combine group with the help of shared markers using MergeMap software (Wu et al., 2011).

2.4 QTL mapping

QTL mapping based on the consensus map was performed using MapQTL v6.0 (Van Ooijen, 2009). The 95% genome-wide significance LOD threshold was determined by Permutation Test with 10,000 permutations. An initial round of regression Interval Mapping was used to scan putative QTLs with a step size of 1 cM. Linkage groups with peaks near or above the significance threshold were chosen for rounds of automatic cofactor selection for Multiple QTL Mapping with a step size of 0.1 cM. For each round of Multiple QTL Mapping, only cofactors within a significant LOD peak were retained, and Multiple QTL Mapping was performed until the QTL position stabilised to determine the true QTLs. QTL intervals with LOD score greater than the thresholds of LG location during the first round of Permutation Test were declared significant. Simultaneously, the position, genetic effects, and percentage of phenotypic variation explained by the QTL were estimated at the peak of significant LOD score. In order to explore the stability of QTLs across the three sites with contrasting environments, QTLs for seven traits were independently detected for each site. QTLs with overlapping genomic regions, detected at different sites for the same trait, were defined as stable QTLs. All figures were drawn using TBtools v1.046 (Chen et al., 2020).

2.5 Transcriptome sequencing and analysis

For each trait (H, D, BD, CC, HC LC, and SG group with H and D sharing a group due to the high correlation between them), six individuals with contrasting phenotypes from 320 progenies of Gonghe population were selected for transcriptome analysis (Table S1). These individuals were selected on the basis of phenotypic measurements made at that site. For example, in the BD group, we selected three individuals with low wood density (0.4431, 0.4374 and 0.4508 g cm -3) and three individuals with high wood density (0.5417, 0.5463 and 0.5506 g cm -3), respectively. Due to the overlap of some individuals between trait groups, we selected a total of 23 individuals with contrasting phenotypes, and information on these individuals is shown in Table S1. The developing xylem tissues of 23 progenies were collected for RNA extraction in mid-July 2018 which is the period for most-active tree growth in Guangdong, China. In a parallel study, 17 of 23 progenies were used to study the effects of gene expression and alternative splicing on wood formation in Eucalyptus (Zhu et al., 2022). The RNA-seq libraries of 23 progenies with contrasting phenotypes were constructed using VAHTS™ Stranded mRNA-seq Library Prep Kit (Vazyme Biotech, China) and were sequenced using the Nextseq-500 platform (Illumina, USA). The raw reads were filtered by removing adaptors and low-quality reads using Trimmomatic v 0.36. Clean reads were mapped to the E. grandis reference genome using HISAT2 (Kim et al., 2015), and StringTie v2.0.4 and Ballgown were used to compute the abundance of the transcripts and identify expressed genes and transcripts in each sample (Pertea et al., 2016).

2.6 Candidate gene identification and functional analysis

Genes located within the QTL confidence interval were extracted according to the positions of the two closest SNPs in the E. grandis reference genome. Then, for each phenotyped trait (H, D, BD, CC, HC, LC, and SG), differentially expressed genes were identified using DESeq2 (Love et al., 2014) with Log2FC ≥ 1 and a false discovery rate (FDR) corrected with P ≤ 0.05. For each trait, genes that were differentially expressed in the progenies and those that were mapped on QTLs of the target trait were assumed to be candidate genes. Gene ontology (GO) enrichment analysis was then applied to the biological processes, molecular functions, and cellular components of the candidate genes using PlantRegMap (Tian et al., 2019) with a threshold of P ≤ 0.05. Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway analysis using Kobas v3.0 (Xie et al., 2011) with a threshold of P ≤ 0.05. Transcription factors (TFs) were predicted for candidate genes using PlantTFDB v4.0 (Jin et al., 2017) with a threshold value of P ≤ 0.05.

3 Results

3.1 Analysis of phenotypic data

Variance component and fixed effect significance, calculated using F statistics based on the method of Kenward and Roger (1997), for each trait, are given in Table S2. There were significant genotype-by-site interactions, except for the HC and LC traits. Only H was significantly different among sites (Table S2). All traits were significantly different among genotypes, as gauged by the Akaike Information Criterion (AIC) (Figures 1B–H; Table S3), where the inclusion of an extra term that results in a lower AIC indicates a significant improvement of the model. Figure S1 gives example histograms showing genotype BLUP values for two traits and the selections that were made. Estimation of pairwise genetic correlations was problematic for most combinations of traits, with estimates of 0.69 between H and CC and – 0.54 between HC and CC being the only combinations resulting in model convergence and that were estimated with acceptable precision (Table S4). This result might be expected, given the small dataset: it is well-established that estimation of trait-trait correlations requires considerable data (e.g. Visscher, 1998). As an alternative, Pearson’ correlations between across-site genotype BLUPs are presented in Figure 1I. Pearson’s correlations between across-sites genotype BLUPs revealed strong, positive correlation between H and D, weak negative correlations between SG and H, D, HC and LC and weak to moderate correlations among most other traits with some correlations involving BD being non-significant (Figure 1I).

3.2 SNP calling and genotyping

GBS generated about 386 Gb of raw data, comprising 1.6 Gb from the E. urophylla, 1.3 Gb from E. tereticornis, and an average of 1.2 Gb from the 320 progenies (Table S5). The GBS seq data are available at NCBI with accession number PRJNA913962. After processing, using process_radtags in Stacks to demultiplex and discard low-quality reads, 7,336,502 clean reads with mean Q20 of 96.6%, mean Q30 of 89.4% and GC content of 41.7% were obtained from the 320 progenies (Table S5). A total of 15,185 high-quality SNPs remained after filtering. Among them, 12,856 (84.7%) SNPs were successfully mapped to the 11 chromosome scaffolds of E. grandis, the remainder being located to as-yet unlocated scaffolds on the E. grandis genome, with the largest number of SNP variants being found on chromosome 6 (Figure 2A). Most SNP variants were observed in intergenic regions (33%), followed by introns (20%), with the fewest in CDS (2%) (Figure 2B).

FIGURE 2
www.frontiersin.org

Figure 2 SNP calling and genetic mapping construction. (A) Distribution of SNP loci on the chromosomes of E. grandis. (B) Genomic position of SNP loci. Syntenic and colinear relationships of each LGs of E. urophylla (C) and E. tereticornis (D) with E. grandis genome sequence. In (E), the curve with the same color represents a similarity match between the chromosomes of E. grandis (e.g., Eg_Chr01) and the corresponding linkage groups of the consensus map (e.g., Co_LG01).

3.3 High-density genetic map construction

Analysis in Lep-Map3 indicated that some markers are effectively redundant, being situated at same locus. This can occur due to close physical location or low recombination rate. Following processing in Lep-Map3, 3686 loci (1615 unique loci) with a total length of 1093.48 cM and a mean interval of 0.68 cM were included in the E. urophylla map, while 2463 loci (1167 unique loci) were present in the E. tereticornis map, with a total length of 1099.18 cM and an average interval of 0.94 cM (Table 1; Table S6). Overall, the E. urophylla map (1093.48 cM) included 448 more loci than the E. tereticornis map (1099.18 cM), but the total genetic map size was very similar (a difference of only 5.70 cM). To further assess the consistency, we BLAST searched the flanking sequences of mapped markers against the E. grandis reference genome (Figures 2C, D). The number of unique loci that could be mapped to the E. grandis reference genome in the E. urophylla and E. tereticornis maps was 1468 (90.9%) and 1064 (91.2%), respectively, and the number of non-syntenic loci are only 47, and 33, respectively, in these two maps (Figures 2C, D; Table S6). The results revealed a high degree of synteny between both the E. urophylla and E. tereticornis genetic maps and the E. grandis reference genome. The consensus map merged using MergeMap contains 2754 loci with a total length of 1359.18 cM, and the shortest and longest linkage groups were Co_LG03 (92.39 cM) and Co_LG06 (185.61 cM), respectively (Table 1 and Figures 3A–B). Compared with the parental maps, the marker density and map quality of the consensus map were further improved (Table 1 and Figure 2E).

TABLE 1
www.frontiersin.org

Table 1 Locus statistics of E. urophylla map, E. tereticornis map and consensus map.

FIGURE 3
www.frontiersin.org

Figure 3 QTL mapping in multiple environments. (A) are the genetic distance in cM. (B) are the distribution of all loci in the consensus map. And the distribution of QTLs of growth and wood properties detected in Gonghe (C), Jijia (D) and Yanxi (E) environments. (F) are stable QTLs. (G) are QTL statistics for each trait at each environment.

3.4 QTL analysis

A total of 108 QTLs associated with growth and wood properties were identified in Gonghe, Jijia and Yanxi sites (Figures 3C–E; Table S7). The LOD scores of these QTLs ranged from 2.32 to 23.27 and explained 2.2 to 23.6% of the phenotypic variation, and 46.3% of the QTLs (50) had positive additive effect (Figure S2, Table S7). Of these, the minimum physical interval of QTL was 214 bp (qBD_YX_2_22.74), the maximum was 4.8 Mb (qCC_YX_3_2.13), and the average was 381.97 kb. Detected QTLs were most frequently associated with LC (23), followed by SG (17), CC (16), D (15), HC (14), BD (12) and H (11), with an average of 15.4 QTLs for each trait (Figure 3G).

3.5 Candidate gene identification and functional analysis

Between one (qBD_YX_2_22.74) and 798 (qBD_GH_10_35.72) genes were identified in the genomic intervals of the aforementioned 108 QTLs (Table S7). Based on RNA-seq, totals of 4119 (H and D), 1258 (BD), 813 (CC), 4355 (HC), 2521 (LC), and 937 (SG) differentially expressed genes were identified in each target trait regions, respectively (Table S8). One to 73 differentially expressed genes were detected in 95 (88.0% of all QTLs) (Table S7). Finally, we identified a total of 1052 candidate genes including some pleiotropic genes that were identified in the QTLs of multiple traits (Figure 4A; Table S9). GO analysis showed that 831 candidate genes had GO annotation and 229 GO terms were significantly enriched (Table S10). The most significant GO terms were response to karrikin (GO:0080167), actin binding (GO:0003779) and cytoskeletal protein binding (GO:0008092) (Figure 4B; Table S10). KEGG analysis revealed 12 significantly enriched pathways, including plant hormone signal transduction, galactose metabolism, and plant-pathogen interaction, suggesting that these pathways may exert important regulatory effects on economic traits in Eucalyptus (Figure 4C; Table S11). Transcription factor prediction indicated that 944 candidate genes significantly enriched 41 TFs from 17 families (Figure 5; Table S12). Among them, a large number of candidate genes were encoding TFs found in the Barley B Recombinant-Basic Pentacysteine (BBR-BPC) and Three-Amino-Acid-Loop-Extension (TALE) family.

FIGURE 4
www.frontiersin.org

Figure 4 Identification and functional analysis of candidate genes for economic traits. (A) Upset plot of the intersection size of candidate genes for each trait. Connected black circles indicate the number of overlapping genes between traits, and unconnected black circles indicate the number of genes unique to a trait. (B) Top 30 most-significant GO terms from analysis of candidate genes. The length of the radial bars indicates the gene count. The CC, MF and BP are cellular component, molecular function and biological process, respectively. (C) shows the top 20 most-significant KEGG pathways. Details of GO terms and KEGG pathways are shown in Tables S10, S11, respectively.

FIGURE 5
www.frontiersin.org

Figure 5 Transcription factor (TF) predicted in candidate genes. Yellow circles are 41 significant TFs and green circles are 944 candidate genes. The color of the line represents the P value. Details of TFs are shown in Table S12.

3.6 QTLs that were stable across sites

Ten QTLs were identified that were stable across at least two sites (Figure 3F; Table S7). Among them, only one QTL, qSG10Stable, was stable at all three sites, which was mapped on chromosome 10 for SG, and explained 23.6%, 18.4% and 19.4% of phenotypic variation in Gonghe, Jijia and Yanxi, respectively (Table S7). Nine QTLs that were stable across two sites were associated with BD (qBD2Stable), D (qD2Stable), SG (qSG3_1Stable, qSG3_2Stable, qSG6_1Stable, qSG6_2Stable) and LC (qLC9_1Stable, qLC9_2Stable, qLC9_3Stable). We also noted that several different QTLs, situated in proximate genomic regions, were detected across different sites. The following examples give the names of QTLs which are a concatenation of the trait code, the site code, the chromosome number and the QTL peak (cM): qCC_YX_3_2.13 and qCC_GH_3_15.36; qLC_JJ_6_67.99, qLC_GH_6_87.31 and qLC_YX_6_98.12. Despite the lack of overlap in the QTL intervals detected across sites, they were taken as potentially stable QTLs. A total of 1173 genes, including 31 candidate genes, were annotated in the physical genomic intervals of the nine stable QTLs (Table S9).

4 Discussion

4.1 High density genetic maps of Eucalyptus were constructed based on GBS technology

In this study, we constructed genetic maps of E. urophylla and E. tereticornis and their consensus maps using large-scale SNP markers generated by GBS. The mean marker spacing of our E. urophylla map, E. tereticornis map and consensus map were 0.66 cM (a total of 1615 SNPs with 1093.48 cM), 0.94 cM (1167 SNPs with 1099.18 cM) and 0.49 cM (2754 SNPs with 1359.18 cM), respectively, representing a considerable increase in marker density and total number of markers mapped, relative to previous maps, constructed using the same populations, employing genomic SSR, EST derived SSR, EST derived CAPS, and DArT markers (Li et al., 2015). These results demonstrate the superiority of current GBS technology in the genetic map construction for Eucalyptus. The marker density of our E. urophylla map and consensus map were slightly lower than that recently reported in E. urophylla (1773 SNPs with 885.9 cM) (Bartholomé et al., 2015) and the consensus map of E. grandis × E. urophylla (2290 markers with 1107.6 cM) (Kullan et al., 2012), however, the E. tereticornis map constructed here is the densest to date (such as, 585 markers with 1241.4 cM in Li et al., 2015; 204 markers with 1023.56 cM in Sumathi et al., 2018). Generally, the density of maps constructed here also compare favourably with recent GBS- or restriction site-associated sequencing (RAD-seq) maps for other forest species. For example, the density of our maps was close to that for Pinus (Hirao et al., 2022) and significantly superior to the Populus maps (Carletti et al., 2016; Xia et al., 2018; Tong et al., 2020).

Some of the differences in resolution of the genetic maps may be explained by the size of the mapping population (Yang et al., 2018b). For example, 1025 F1 progenies were employed in the previous E. urophylla map (Bartholomé et al., 2015), compared with 320 F1 progenies in ours and 130 F1 offspring in Pinus map (Hirao et al., 2022), and 131 to 300 F1 progenies in Populus maps (Carletti et al., 2016; Xia et al., 2018; Tong et al., 2020). Although preserving large mapping populations is difficult and rarely implemented in forest trees, it may contribute to the accuracy and reliability of genetic linkage mapping and subsequent QTL mapping.

It is also noted that larger and more-complex genomes make SNP calling and genetic mapping more difficult. For example, the genome of Pinus is very large – P. taeda is around 20 Gb - and complex, including repetitive sequences, transposable elements, and gene duplication (Zimin et al., 2014). In contrast, we used a simplified strategy for SNP calling in Eucalyptus, which possesses a much smaller genome, using a combination of double digests. The higher resolution of the previous E. urophylla map may have benefited from the greater number of genome-wide SNP variants detected by whole-genome resequencing technology (Bartholomé et al., 2015). Similarly, ultra-high-resolution genetic linkage maps were reported in several crops based on whole-genome resequencing (Luo et al., 2020; He et al., 2021; Kong et al., 2022). It should be noted that the map resolution is only one of the possible indicators of map quality, considering the differences in genome size and complexity among different genera, the differences in algorithms and functions of different mapping software, and the effects of biased separation markers.

While GBS technology provides very numerous markers, enabling construction of relatively high-density genetic maps, due to the markers being unevenly distributed within the genome (Figure 2), regions of low marker density still exist. For example, we detected a gap of 16.51 cM in Et_LG09 91.37–107.88 cM and a gap of 8.87 cM in Co_LG09 5.51–14.38 cM (Tables 1 and S6). These low density regions may be caused by low polymorphism of genetic variation (Kullan et al., 2012). Alternatively, the efficiency of restriction enzyme digestion selected in GBS varies in species (Wang et al., 2017), which may result in low genome coverage. Higher depth, simplified genome sequencing complemented by whole genome resequencing with broader coverage may be a strategy that provides better uniformity and higher density. In addition, the selection of appropriate restriction enzymes can also improve the SNP calling efficiency. It is worth remarking that the type II restriction enzymes (e.g., MspI-MseI) used in this study tend to result in low coverage of the methylation regions of the genome (Hashimoto et al., 2007). Whether or not the low marker density regions in our genetic map correspond to abundant methylation regions deserves further investigation. In future, high-depth coverage and analysis of all genomic regions might be done through development of methylation markers that can be combined with the aforementioned strategies of high depth, simplified GBS and whole-genome resequencing. Overall, our results confirm that the low-cost of GBS technology provides tens of thousands of genomic loci, significantly improving the genetic map resolution of Eucalyptus.

4.2 QTL mapping and candidate gene identification of growth and wood properties

A total of 108 QTLs associated with economic traits were identified at three sites, and the number of QTLs for most traits was higher than those identified in previous studies (Freeman et al., 2009; Gion et al., 2011; Li et al., 2015; Sumathi et al., 2018). Taking wood density as an example, Freeman et al. (2009) detected three QTLs on three LGs, Li et al. (2015) detected four QTLs on four LGs, and we detected 12 QTLs on six LGs. Most of the 12 QTLs were not identified in previous studies, suggesting that high-density genetic mapping reveals more QTLs including some that are novel. Moreover, in further comparison, the phenotypic variance explained (PVE) of the previous wood density QTLs ranged from 9.5 to 15.5 (Freeman et al., 2009; Li et al., 2015), while the average PVE of our 12 QTLs was only 4.9, with a maximum of 11.2 (qBD_YX_2_106.18). This indicates that that the economic traits of E. urophylla × tereticornis may be regulated by numerous QTLs, each of small effect, which accords with the assumption that these are quantitative traits.

QTL mapping and RNA-seq are two different approaches that focus on the identification of trait-related genes at different regulatory stages. Target trait genes identified by QTL mapping tend to be more involved in upstream molecular regulation (He et al., 2021). In contrast, due to the transient nature of RNA-seq, more of the genes that are involved in regulation tend to be identified (Wang et al., 2018). In addition, some studies have used post-metabolic products to infer the genes involved in regulation (Galpaz et al., 2018; Peng et al., 2022). Our results showed that 88.0% of the QTLs were involved in transcriptional regulation (Table S7), indicating that the results of QTL mapping and RNA-seq can be efficiently inter-validated.

Examination of known gene function indicates that many of the 1052 candidate genes have plausible links to the regulation of economic traits in E. urophylla × tereticornis (Table S9). These candidate genes are mainly associated with sugar metabolism, signal transduction, morphogenesis and response, as well as several known wood biosynthetic pathways (Zhang et al., 2016). We found that many candidate genes are involved in the encoding of TFs (Figure 5). For example, TALE, a class of homologous structural domain proteins, has been shown to play an important role in the secondary growth of plant cell walls and the development of root and stem (Magnani and Hake, 2008). In this study, 479 candidate genes encoding TALEs were identified.

In addition to genes encoding TALEs, some TFs were also found. MYB is a class of plant-specific TFs in the plant genome. In Arabidopsis thaliana, MYB46 and MYB83 have been confirmed as secondary master switches regulating secondary wall formation, and overexpression of both genes results in ectopic secondary wall deposition (Zhong et al., 2007; McCarthy et al., 2009). In Populus, more than 15 MYBs have been reported to be involved in secondary wall formation, of which six were confirmed to be secondary master switches (Li et al., 2020a). In the present study, 105 candidate genes encoding three MYBs were also detected (Table S12). We hypothesize that these candidate genes may have a switch function similar to that in Populus, regulating the wood formation process in Eucalyptus. In addition, TFs of the NAC, Lateral Organ Boundaries Domain (LBD), DNA binding with one finger (Dof) and Basic Helix-Loop-Helix (bHLH) and other families have been reported to be involved in the growth and wood formation process (Yordanov et al., 2010; Hussey et al., 2011; Xiang et al., 2019; Sun et al., 2021). Our results suggest that these candidate genes may have important effects on economic traits by encoding TFs, and those genomic resources for further research will accelerate breeding for economic traits in Eucalyptus and also help to reveal the genetic basis of wood formation.

4.3 Stable QTLs

In the context of a considerable number of Eucalyptus clones that have been successfully used in production forestry, screening for stable QTLs among specific hybrid progeny is a realistic strategy to guide practical breeding. Therefore, we used clonally reproduced full-sibs of a hybrid deployed across sites to carry out QTL stability studies. This is an important difference between the present and previous studies (Freeman et al., 2011; Freeman et al., 2013). The use of clonal replication within and across sites reduces the effect of genetic variation on phenotypic variation and thus allows better partitioning of environmental variance. Among 108 QTLs, the vast majority were not stable across the three study sites, which is consistent with previous findings in E. globulus (Freeman et al., 2013), C. japonica (Mori et al., 2019; Mori et al., 2020), P. taeda (Westbrook et al., 2015), and some crops (Raihan et al., 2016; He et al., 2021). These studies suggested that quantitative traits are influenced by interactions between genetic and environmental factors, resulting in the identification of specific QTLs in specific environments only.

Of the ten QTLs detected at more than two sites (Figure 3F and Table S7), only qSG10Stable was stable across all three sites, mapped between positions 29,512,351 and 30,943,933 bp of chromosome 10. It is interesting to note that the two significant SG loci identified on chromosome 10 (30,915,259 bp) in E. globulus (Cappa et al., 2013) were located exactly in the qSG10Stable interval. Similarly, Mizrachi et al. (2017) identified two eQTLs associated with LC on chromosome 10, from which a large number of genes associated with the lignin biosynthetic process were identified. In qSG10Stable, we identified two candidate genes (Eucgr.J02456, Eucgr.J02459) encoding glycosyl hydrolase family 38 proteins (Table S9). Previously, Wei et al. (2017) integrated RNA-seq and GWAS analysis and identified one candidate gene encoding a glycosyl hydrolase that regulates the LC of Brassica napus. This evidence suggests a strong association signal between qSG10Stable and lignin biosynthesis.

We found that qSG10Stable explained the most phenotypic variation at all three sites (PVE: 23.6% in Gonghe, 18.4% in Jijia, 19.4% in Yanxi), suggesting that QTLs with strong effects have higher likelihood of stability across environments, a finding supported by previous studies (Wan et al., 2005; Raihan et al., 2016; He et al., 2021). However, the PVE of stable QTLs may change in different environments due to the interaction between QTL and environment, for example, the PVEs of the other nine partially stable QTLs (i.e. those stable across two sites only) are highly variable. We found only one partially stable QTL, qD2Stable, associated with growth traits. This might be expected, as growth traits tend to be subject to stronger QTL-by environment interactions than wood properties (Bartholomé et al., 2013; Freeman et al., 2013; Mori et al., 2019).

Climatic factors may have important effects on the QTL stability across environments. For instance, in a multi-environment (two relatively drier and one wetter regions) QTL study of C. japonica, two stable QTLs were detected only in the two drier regions and not in the wetter region, and their stability was considered to be related to limited water availability during the growing season (Mori et al., 2019). Our study sites included two subtropical monsoon climates, Gonghe and Yanxi, and one tropical maritime climate, Jijia, which are climatically significantly different (Yang et al., 2018a). It would therefore seem likely that studies involving more sites or sites with widely varying climates would identify even fewer QTLs that are stable across all sites. However, such studies, although expensive to establish, may allow us to better assess the nature of QTL × environment interactions and potentially identify QTLs that are at least partially stable across a predictable subset of site types. Moreover, as most of these stable QTLs are associated with pulp traits, they are valuable for future breeding of E. urophylla × tereticornis hybrids.

Data availability statement

The GBS seq data has been deposited at NCBI with accession number PRJNA913962.

Author contributions

XZ: performed molecular lab work and wrote the manuscript with assistance from all other authors; QW, CZ, DB: phenotype traits measurement and analysis; QW, CZ, FL: maintained the field trial and collected the leaf samples; XZ, DB, HZ, PW, FL: analyzed the data; DB: review and editing the manuscript; FL: conceived and designed the project, review and editing the manuscript. All authors read and approved final manuscript.

Funding

This work was financially supported by the Fundamental Research Funds of Chinese Academy of Forestry (CAFYBB2021ZA001; CAFYBB2021SY001; CAFYBB2021QD001-3).

Acknowledgments

We thank Dr. Siming Gan and his colleagues for their coordination of the research project, and for the management and cultivation of clonal materials.

Conflict of interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher’s note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

Supplementary material

The Supplementary Material for this article can be found online at: https://www.frontiersin.org/articles/10.3389/fpls.2022.1099705/full#supplementary-material

Supplementary Figure 1 | Histograms showing best linear unbiased prediction deviations (across-sites genotype values) for inter-specific hybrid eucalypt full-sib progeny from across-sites mixed model analyses for HC and SG traits. Red bars indicate individuals selected on the basis of their above- or below-average genetic values. Some genotypes included more than one selected individual.

Supplementary Figure 2 | LOD scores for linkage groups of consensus map for each trait in three environments. The LOD thresholds of QTL are given in Table S7.

Supplementary Table 1 | Twenty-three individuals with contrasting phenotypes for RNA-seq.

Supplementary Table 2 | Analysis of linear mixed models. Random effects by trait for block-replicate and row (both within site), genotype, genotype-by-site interaction (GxS) and heterogeneous site residuals and fixed-effect F statistic, denominator degrees of freedom (d.d.f.) and F probability for the fixed site term. For each variance component estimate, Z (the component divided by it standard error) is given. Informally, Z>2 indicates a term that is likely to be gauged significant using the AIC. *Con. Variance component was constrained to zero with zero standard error.

Supplementary Table 3 | Akaike Information Criterion and grand means from mixed model analyses (means net of block, column and genotype BLUP effects) for each trait.

Supplementary Table 4 | Bivariate genetic correlations between pairs of traits with correlation estimates (lower triangle - B denotes a boundary estimate) and Z (upper triangle - the estimated correlation coefficient divided by its standard error– given where estimable).

Supplementary Table 5 | Data statistics of GBS sequencing.

Supplementary Table 6 | Marker information in the genetic maps.

Supplementary Table 7 | Information of 108 QTLs and ten stable QTLs for economic traits.

Supplementary Table 8 | Differentially expressed genes identified in each trait group by RNA-seq.

Supplementary Table 9 | Information of 1052 candidate genes and genes in stable QTLs

Supplementary Table 10 | GO enrichment of candidate genes.

Supplementary Table 11 | KEGG enrichment of candidate genes.

Supplementary Table 12 | Transcription factor prediction of candidate genes.

References

Baltunis, B. S., Wu, H. X., Powell, M. B. (2007). Inheritance of density, microfibril angle, and modulus of elasticity in juvenile wood of pinus radiata at two locations in Australia. Can. J. For. Res. 37, 2164–2174. doi: 10.1139/X07-061

CrossRef Full Text | Google Scholar

Bartholomé, J., Mandrou, E., Mabiala, A., Jenkins, J., Nabihoudine, I., Klopp, C., et al. (2015). High-resolution genetic maps of Eucalyptus improve Eucalyptus grandis genome assembly. New Phytol. 206, 1283–1296. doi: 10.1111/nph.13150

PubMed Abstract | CrossRef Full Text | Google Scholar

Bartholomé, J., Salmon, F., Vigneron, P., Bouvet, J.-M., Plomion, C., Gion, J.-M. (2013). Plasticity of primary and secondary growth dynamics in Eucalyptus hybrids: a quantitative genetics and QTL mapping perspective. BMC Plant Biol. 13, 120. doi: 10.1186/1471-2229-13-120

PubMed Abstract | CrossRef Full Text | Google Scholar

Bolger, A. M., Lohse, M., Usadel, B. (2014). Trimmomatic: a flexible trimmer for illumina sequence data. Bioinformatics 30, 2114–2120. doi: 10.1093/bioinformatics/btu170

PubMed Abstract | CrossRef Full Text | Google Scholar

Brawner, J. T., Lee, D. J., Meder, R., Almeida, A. C., Dieters, M. J. (2013). Classifying genotype by environment interactions for targeted germplasm deployment with a focus on Eucalyptus. Euphytica 191, 403–414. doi: 10.1007/s10681-013-0892-4

CrossRef Full Text | Google Scholar

Brenton, v. T.D., Cameron, J., David, B. L., Lindenmayer, D. B., Sam, C. B. (2019). Pervasive admixture between eucalypt species has consequences for conservation and assisted migration. Evolutionary Appl. 12, 845–860. doi: 10.1111/eva.12761

CrossRef Full Text | Google Scholar

Bush, D., Kain, D., Kanowski, P., Matheson, C. (2015). Genetic parameter estimates informed by a marker-based pedigree: a case study with eucalyptus cladocalyx in southern Australia. Tree Genet. Genomes 11, 798. doi: 10.1007/s11295-014-0798-x

CrossRef Full Text | Google Scholar

Butler, D. G., Cullis, B. R., Gilmour, A. R., Gogel, B. J., Thompson, R. (2018). ASReml-r reference manual version 4 (Hemel Hempstead, UK: VSN International Ltd).

Google Scholar

Cappa, E. P., El-Kassaby, Y. A., Garcia, M. N., Acuña, C., Borralho, N. M. G., Grattapaglia, D., et al. (2013). Impacts of population structure and analytical models in genome-wide association studies of complex traits in forest trees: A case study in Eucalyptus globulus. PloS One 8, e81267. doi: 10.1371/journal.pone.0081267

PubMed Abstract | CrossRef Full Text | Google Scholar

Carletti, G., Carra, A., Allegro, G., Vietto, L., Desiderio, F., Bagnaresi, P., et al. (2016). QTLs for woolly poplar aphid (Phloeomyzus passerinii l.) resistance detected in an inter-specific Populus deltoides × P. nigra mapping population. PloS One 11, e0152569. doi: 10.1371/journal.pone.0152569

PubMed Abstract | CrossRef Full Text | Google Scholar

Chen, C., Chen, H., Zhang, Y., Thomas, H. R., Frank, M. H., He, Y., et al. (2020). TBtools: An integrative toolkit developed for interactive analyses of big biological data. Mol. Plant 13, 1194–1202. doi: 10.1016/j.molp.2020.06.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Cingolani, P., Platts, A., Wang, L. L., Coon, M., Nguyen, T., Wang, L., et al. (2012). A program for annotating and predicting the effects of single nucleotide polymorphisms, SnpEff. null 6, 80–92. doi: 10.4161/fly.19695

CrossRef Full Text | Google Scholar

de Araujo, M. J., de Paula, R. C., Campoe, O. C., Carneiro, R. L. (2019). Adaptability and stability of eucalypt clones at different ages across environmental gradients in Brazil. For. Ecol. Manage. 454, 117631. doi: 10.1016/j.foreco.2019.117631

CrossRef Full Text | Google Scholar

Downes, G., Hudson, I., Raymond, C., Dean, G., Michell, A., Schimleck, L., et al. (1997). Sampling plantation eucalypts for wood and fibre properties (Melbourne: CSIRO Publishing).

Google Scholar

Elshire, R. J., Glaubitz, J. C., Sun, Q., Poland, J. A., Kawamoto, K., Buckler, E. S., et al. (2011). A robust, simple genotyping-by-Sequencing (GBS) approach for high diversity species. PloS One 6, e19379. doi: 10.1371/journal.pone.0019379

PubMed Abstract | CrossRef Full Text | Google Scholar

Freeman, J. S., Potts, B. M., Downes, G. M., Pilbeam, D., Thavamanikumar, S., Vaillancourt, R. E. (2013). Stability of quantitative trait loci for growth and wood properties across multiple pedigrees and environments in Eucalyptus globulus. New Phytol. 198, 1121–1134. doi: 10.1111/nph.12237

PubMed Abstract | CrossRef Full Text | Google Scholar

Freeman, J. S., Potts, B. M., Downes, G. M., Thavamanikumar, S., Pilbeam, D. J., Hudson, C. J., et al. (2011). QTL analysis for growth and wood properties across multiple pedigrees and sites in eucalyptus globulus. BMC Proc. 5, O8. doi: 10.1186/1753-6561-5-S7-O8

PubMed Abstract | CrossRef Full Text | Google Scholar

Freeman, J. S., Whittock, S. P., Potts, B. M., Vaillancourt, R. E. (2009). QTL influencing growth and wood properties in Eucalyptus globulus. Tree Genet. Genomes 5, 713–722. doi: 10.1007/s11295-009-0222-0

CrossRef Full Text | Google Scholar

Galpaz, N., Gonda, I., Shem-Tov, D., Barad, O., Tzuri, G., Lev, S., et al. (2018). Deciphering genetic factors that determine melon fruit-quality traits using RNA -seq-based high-resolution QTL and eQTL mapping. Plant J. 94, 169–191. doi: 10.1111/tpj.13838

PubMed Abstract | CrossRef Full Text | Google Scholar

Gan, S., Shi, J., Li, M., Wu, K., Wu, J., Bai, J. (2003). Moderate-density molecular maps of Eucalyptus urophylla s. t. Blake and E. tereticornis smith genomes based on RAPD markers. Genetica 118, 59–67. doi: 10.1023/A:1022966018079

PubMed Abstract | CrossRef Full Text | Google Scholar

Gion, J.-M., Carouché, A., Deweer, S., Bedon, F., Pichavant, F., Charpentier, J.-P., et al. (2011). Comprehensive genetic dissection of wood properties in a widely-grown tropical tree: Eucalyptus. BMC Genomics 12, 301. doi: 10.1186/1471-2164-12-301

PubMed Abstract | CrossRef Full Text | Google Scholar

Gong, X., Yang, A., Wu, Z., Chen, C., Li, H., Liu, Q., et al. (2021). Employing genome-wide SNP discovery to characterize the genetic diversity in Cinnamomum camphora using genotyping by sequencing. Forests 12, 1511. doi: 10.3390/f12111511

CrossRef Full Text | Google Scholar

Grattapaglia, D., Kirst, M. (2008). Eucalyptus applied genomics: from gene sequences to breeding tools. New Phytol. 179, 911–929. doi: 10.1111/j.1469-8137.2008.02503.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Hall, D., Hallingbäck, H. R., Wu, H. X. (2016). Estimation of number and size of QTL effects in forest tree traits. Tree Genet. Genomes 12, 110. doi: 10.1007/s11295-016-1073-0

CrossRef Full Text | Google Scholar

Hashimoto, K., Kokubun, S., Itoi, E., Roach, H. I. (2007). Improved quantification of DNA methylation using methylation-sensitive restriction enzymes and real-time PCR. Epigenetics 2, 86–91. doi: 10.4161/epi.2.2.4203

PubMed Abstract | CrossRef Full Text | Google Scholar

He, Q., Zhi, H., Tang, S., Xing, L., Wang, S., Wang, H., et al. (2021). QTL mapping for foxtail millet plant height in multi-environment using an ultra-high density bin map. Theor. Appl. Genet. 134, 557–572. doi: 10.1007/s00122-020-03714-w

PubMed Abstract | CrossRef Full Text | Google Scholar

Hirao, T., Matsunaga, K., Shirasawa, K. (2022). Quantitative trait loci analysis based on high-density mapping of single-nucleotide polymorphisms by genotyping-by-Sequencing against pine wilt disease in Japanese black pine (Pinus thunbergii). Front. Plant Sci. 13. doi: 10.3389/fpls.2022.850660

PubMed Abstract | CrossRef Full Text | Google Scholar

Hussey, S. G., Mizrachi, E., Spokevicius, A. V., Bossinger, G., Berger, D. K., Myburg, A. A. (2011). SND2, a NAC transcription factor gene, regulates genes involved in secondary cell wall development in arabidopsis fibres and increases fibre cell area in Eucalyptus. BMC Plant Biol. 11, 173. doi: 10.1186/1471-2229-11-173

PubMed Abstract | CrossRef Full Text | Google Scholar

Jin, J., Tian, F., Yang, D.-C., Meng, Y.-Q., Kong, L., Luo, J., et al. (2017). PlantTFDB 4.0: toward a central hub for transcription factors and regulatory interactions in plants. Nucleic Acids Res. 45, D1040–D1045. doi: 10.1093/nar/gkw982

PubMed Abstract | CrossRef Full Text | Google Scholar

Kenward, M. G., Roger, J. H. (1997). Small sample inference for fixed effects from restricted maximum likelihood. Biometrics 53, 983–997. doi: 10.2307/2533558

PubMed Abstract | CrossRef Full Text | Google Scholar

Kim, D., Langmead, B., Salzberg, S. L. (2015). HISAT: a fast spliced aligner with low memory requirements. Nat. Methods 12, 357–360. doi: 10.1038/nmeth.3317

PubMed Abstract | CrossRef Full Text | Google Scholar

Klápště, J., Ashby, R. L., Telfer, E. J., Graham, N. J., Dungey, H. S., Brauning, R., et al. (2021). The use of "Genotyping-by-Sequencing" to recover shared genealogy in genetically diverse Eucalyptus populations. Forests 12, 904. doi: 10.3390/f12070904

CrossRef Full Text | Google Scholar

Kong, W., Deng, X., Yang, J., Zhang, C., Sun, T., Ji, W., et al. (2022). High-resolution bin-based linkage mapping uncovers the genetic architecture and heterosis-related loci of plant height in indica–japonica derived populations. Plant J. 110, 814–827. doi: 10.1111/tpj.15705

PubMed Abstract | CrossRef Full Text | Google Scholar

Kullan, A. R. K., van Dyk, M. M., Jones, N., Kanzler, A., Bayley, A., Myburg, A. A. (2012). High-density genetic linkage maps with over 2,400 sequence-anchored DArT markers for genetic dissection in an F2 pseudo-backcross of Eucalyptus grandis × E. urophylla. Tree Genet. Genomes 8, 163–175. doi: 10.1007/s11295-011-0430-2

CrossRef Full Text | Google Scholar

Ladiges, P. Y., Udovicic, F., Nelson, G. (2003). Australian Biogeographical connections and the phylogeny of large genera in the plant family myrtaceae: Australian biogeographical connections. J. Biogeography 30, 989–998. doi: 10.1046/j.1365-2699.2003.00881.x

CrossRef Full Text | Google Scholar

Laoué, J., Depardieu, C., Gérardi, S., Lamothe, M., Bomal, C., Azaiez, A., et al. (2021). Combining QTL mapping and transcriptomics to decipher the genetic architecture of phenolic compounds metabolism in the conifer white spruce. Front. Plant Sci. 12. doi: 10.3389/fpls.2021.675108

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, H., Guo, X., Liu, Y., Lu, H. (2020a). The molecular mechanism in secondary wall deposition and programmed cell death of wood formation (in Chinese). Sci. Sin. Vitae 50, 123–135. doi: 10.1360/SSV-2019-0133

CrossRef Full Text | Google Scholar

Li, W., Wu, H., Li, X., Chen, Y., Yin, T. (2020b). Fine mapping of the sex locus in Salix triandra confirms a consistent sex determination mechanism in genus salix. Hortic. Res. 7, 64. doi: 10.1038/s41438-020-0289-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, F., Zhou, C., Weng, Q., Li, M., Yu, X., Guo, Y., et al. (2015). Comparative genomics analyses reveal extensive chromosome colinearity and novel quantitative trait loci in Eucalyptus. PloS One 10, e0145144. doi: 10.1371/journal.pone.0145144

PubMed Abstract | CrossRef Full Text | Google Scholar

Love, M. I., Huber, W., Anders, S. (2014). Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biol. 15, 550. doi: 10.1186/s13059-014-0550-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Luo, X., Xu, L., Wang, Y., Dong, J., Chen, Y., Tang, M., et al. (2020). An ultra-high-density genetic map provides insights into genome synteny, recombination landscape and taproot skin colour in radish (Raphanus sativus l.). Plant Biotechnol. J. 18, 274–286. doi: 10.1111/pbi.13195

PubMed Abstract | CrossRef Full Text | Google Scholar

Magnani, E., Hake, S. (2008). KNOX lost the OX: The Arabidopsis KNATM gene defines a novel class of KNOX transcriptional regulators missing the homeodomain. Plant Cell 20, 875–887. doi: 10.1105/tpc.108.058495

PubMed Abstract | CrossRef Full Text | Google Scholar

Malan, F. S., Verryn, S. D. (1996). Effect of genotype-by-Environment interaction on the wood properties and qualities of four-year-old Eucalyptus grandis and E. grandis hybrids. South Afr. Forestry J. 176, 47–53. doi: 10.1080/00382167.1996.9629709

CrossRef Full Text | Google Scholar

Maliepaard, C., Jansen, J., Van Ooijen, J. W. (1997). Linkage analysis in a full-sib family of an outbreeding plant species: overview and consequences for applications. Genet. Res. 70, 237–250. doi: 10.1017/S0016672397003005

CrossRef Full Text | Google Scholar

McCarthy, R. L., Zhong, R., Ye, Z.-H. (2009). MYB83 is a direct target of SND1 and acts redundantly with MYB46 in the regulation of secondary cell wall biosynthesis in Arabidopsis. Plant Cell Physiol. 50, 1950–1964. doi: 10.1093/pcp/pcp139

PubMed Abstract | CrossRef Full Text | Google Scholar

Mizrachi, E., Verbeke, L., Christie, N., Fierro, A. C., Mansfield, S. D., Davis, M. F., et al. (2017). Network-based integration of systems genetics data reveals pathways associated with lignocellulosic biomass accumulation and processing. Proc. Natl. Acad. Sci. U.S.A. 114, 1195–1200. doi: 10.1073/pnas.1620119114

PubMed Abstract | CrossRef Full Text | Google Scholar

Mori, H., Ueno, S., Ujino-Ihara, T., Fujiwara, T., Yamashita, K., Kanetani, S., et al. (2019). Mapping quantitative trait loci for growth and wood property traits in Cryptomeria japonica across multiple environments. Tree Genet. Genomes 15, 43. doi: 10.1007/s11295-019-1346-5

CrossRef Full Text | Google Scholar

Mori, H., Yamashita, K., Saiki, S.-T., Matsumoto, A., Ujino-Ihara, T. (2020). Climate sensitivity of Cryptomeria japonica in two contrasting environments: Perspectives from QTL mapping. PloS One 15, e0228278. doi: 10.1371/journal.pone.0228278

PubMed Abstract | CrossRef Full Text | Google Scholar

Myburg, A. A., Grattapaglia, D., Tuskan, G. A., Hellsten, U., Hayes, R. D., Grimwood, J., et al. (2014). The genome of Eucalyptus grandis. Nature 510, 356–362. doi: 10.1038/nature13308

PubMed Abstract | CrossRef Full Text | Google Scholar

Nakahama, K., Urata, N., Shinya, T., Hayashi, K., Nanto, K., Rosa, A. C., et al. (2018). RNA-Seq analysis of lignocellulose-related genes in hybrid Eucalyptus with contrasting wood basic density. BMC Plant Biol. 18, 156. doi: 10.1186/s12870-018-1371-9

PubMed Abstract | CrossRef Full Text | Google Scholar

Pavy, N., Lamothe, M., Pelgas, B., Gagnon, F., Birol, I., Bohlmann, J., et al. (2017). A high-resolution reference genetic map positioning 8.8 K genes for the conifer white spruce: structural genomics implications and correspondence with physical distance. Plant J. 90, 189–203. doi: 10.1111/tpj.13478

PubMed Abstract | CrossRef Full Text | Google Scholar

Peng, Z., Zhao, C., Li, S., Guo, Y., Xu, H., Hu, G., et al. (2022). Integration of genomics, transcriptomics and metabolomics identifies candidate loci underlying fruit weight in loquat. Horticult. Res. 9, uhac037. doi: 10.1093/hr/uhac037

CrossRef Full Text | Google Scholar

Pertea, M., Kim, D., Pertea, G. M., Leek, J. T., Salzberg, S. L. (2016). Transcript-level expression analysis of RNA-seq experiments with HISAT, StringTie and ballgown. Nat. Protoc. 11, 1650–1667. doi: 10.1038/nprot.2016.095

PubMed Abstract | CrossRef Full Text | Google Scholar

Poland, J. A., Brown, P. J., Sorrells, M. E., Jannink, J.-L. (2012). Development of high-density genetic maps for barley and wheat using a novel two-enzyme genotyping-by-Sequencing approach. PloS One 7, e32253. doi: 10.1371/journal.pone.0032253

PubMed Abstract | CrossRef Full Text | Google Scholar

Rae, A. M., Pinel, M. P. C., Bastien, C., Sabatti, M., Street, N. R., Tucker, J., et al. (2007). QTL for yield in bioenergy Populus: identifying G×E interactions from growth at three contrasting sites. Tree Genet. Genomes 4, 97–112. doi: 10.1007/s11295-007-0091-3

CrossRef Full Text | Google Scholar

Raihan, M. S., Liu, J., Huang, J., Guo, H., Pan, Q., Yan, J. (2016). Multi-environment QTL analysis of grain morphology traits and fine mapping of a kernel-width QTL in Zheng58 × SK maize population. Theor. Appl. Genet. 129, 1465–1477. doi: 10.1007/s00122-016-2717-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Rastas, P., Paulin, L., Hanski, I., Lehtonen, R., Auvinen, P. (2013). Lep-MAP: fast and accurate linkage map construction for large SNP datasets. Bioinformatics 29, 3128–3134. doi: 10.1093/bioinformatics/btt563

PubMed Abstract | CrossRef Full Text | Google Scholar

Rochette, N. C., Rivera-Colón, A. G., Catchen, J. M. (2019). Stacks 2: Analytical methods for paired-end sequencing improve RADseq-based population genomics. Mol. Ecol. 28, 4737–4754. doi: 10.1111/mec.15253

PubMed Abstract | CrossRef Full Text | Google Scholar

Shang, L., Liang, Q., Wang, Y., Wang, X., Wang, K., Abduweli, A., et al. (2015). Identification of stable QTLs controlling fiber traits properties in multi-environment using recombinant inbred lines in upland cotton (Gossypium hirsutum l.). Euphytica 205, 877–888. doi: 10.1007/s10681-015-1434-z

CrossRef Full Text | Google Scholar

Sumathi, M., Bachpai, V. K. W., Deeparaj, B., Mayavel, A., Dasgupta, M. G., Nagarajan, B., et al. (2018). Quantitative trait loci mapping for stomatal traits in interspecific hybrids of Eucalyptus. J. Genet. 97, 323–329. doi: 10.1007/s12041-018-0896-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Sun, J., Guo, Y., Li, S., Zhou, C., Chiang, V., Li, W. (2021). A functional study of bHLH106 transcription factor based on CRISPR / Cas9 in Populus trichocarpa. J. Nanjing Forestry Univ. (Natural Sci. Edition) 45 (6), 15–23. doi: 10.12302/j.issn.1000-2006.202107031

CrossRef Full Text | Google Scholar

Thumma, B. R., Southerton, S. G., Bell, J. C., Owen, J. V., Henery, M. L., Moran, G. F. (2010). Quantitative trait locus (QTL) analysis of wood quality traits in Eucalyptus nitens. Tree Genet. Genomes 6, 305–317. doi: 10.1007/s11295-009-0250-9

CrossRef Full Text | Google Scholar

Tian, F., Yang, D.-C., Meng, Y.-Q., Jin, J., Gao, G. (2019). PlantRegMap: charting functional regulatory maps in plants. Nucleic Acids Res. 48, D1104–D1113. doi: 10.1093/nar/gkz1020

CrossRef Full Text | Google Scholar

Tong, C., Yao, D., Wu, H., Chen, Y., Yang, W., Zhao, W. (2020). High-quality SNP linkage maps improved QTL mapping and genome assembly in Populus. J. Heredity 111, 515–530. doi: 10.1093/jhered/esaa039

CrossRef Full Text | Google Scholar

Vaillant, A., Honvault, A., Bocs, S., Summo, M., Makouanzi, G., Vigneron, P., et al. (2018). Genetic effect in leaf and xylem transcriptome variations among Eucalyptus urophylla × grandis hybrids in field conditions. Silvae Genetica 67, 57–65. doi: 10.2478/sg-2018-0008

CrossRef Full Text | Google Scholar

Van Ooijen, J. (2006). JoinMap 4.0, software for the calculation of genetic linkage maps in experimental populations. (Kyazma B.V, Wageningen).

Google Scholar

Van Ooijen, J. (2009). MapQTL 6.0, software for the mapping of quantitative trait loci in experimental populations of dihaploid species. Netherlands: Wageningen: Kyazma BV.

Google Scholar

Visscher, P. M. (1998). On the sampling variance of intraclass correlations and genetic correlations. Genetics 149, 1605–1614. doi: 10.1093/genetics/149.3.1605

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, Y., Cao, X., Zhao, Y., Fei, J., Hu, X., Li, N. (2017). Optimized double-digest genotyping by sequencing (ddGBS) method with high-density SNP markers and high genotyping accuracy for chickens. PloS One 12, e0179073. doi: 10.1371/journal.pone.0179073

PubMed Abstract | CrossRef Full Text | Google Scholar

Wang, T., Hou, L., Jian, H., Di, F., Li, J., Liu, L. (2018). Combined QTL mapping, physiological and transcriptomic analyses to identify candidate genes involved in Brassica napus seed aging. Mol. Genet. Genomics 293, 1421–1435. doi: 10.1007/s00438-018-1468-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Wan, X. Y., Wan, J. M., Weng, J. F., Jiang, L., Bi, J. C., Wang, C. M., et al. (2005). Stability of QTLs for rice grain dimension and endosperm chalkiness characteristics across eight environments. Theor. Appl. Genet. 110, 1334–1346. doi: 10.1007/s00122-005-1976-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Wei, L., Jian, H., Lu, K., Yin, N., Wang, J., Duan, X., et al. (2017). Genetic and transcriptomic analyses of lignin- and lodging-related traits in Brassica napus. Theor. Appl. Genet. 130, 1961–1973. doi: 10.1007/s00122-017-2937-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Westbrook, J. W., Walker, A. R., Neves, L. G., Munoz, P., Resende, M. F. R., Neale, D. B., et al. (2015). Discovering candidate genes that regulate resin canal number in Pinus taeda stems by integrating genetic analysis across environments, ages, and populations. New Phytol. 205, 627–641. doi: 10.1111/nph.13074

PubMed Abstract | CrossRef Full Text | Google Scholar

Wu, Y., Close, T. J., Lonardi, S. (2011). Accurate construction of consensus genetic maps via integer linear programming. Ieee-Acm Transact Comput. Biol. Bioinform. 8 (2), 381–394. doi: 10.1109/TCBB.2010.35

CrossRef Full Text | Google Scholar

Wu, R. L., Ma, C. X., Painter, I., Zeng, Z. B. (2002). Simultaneous maximum likelihood estimation of linkage and linkage phases in outcrossing species. Theor. Popul Biol. 61, 349–363. doi: 10.1006/tpbi.2002.1577

PubMed Abstract | CrossRef Full Text | Google Scholar

Xiang, W.-B., Li, W.-F., Zhang, S.-G., Qi, L.-W. (2019). Transcriptome-wide analysis to dissect the transcription factors orchestrating the phase change from vegetative to reproductive development in Larix kaempferi. Tree Genet. Genomes 15, 68. doi: 10.1007/s11295-019-1376-z

CrossRef Full Text | Google Scholar

Xia, W., Xiao, Z., Cao, P., Zhang, Y., Du, K., Wang, N. (2018). Construction of a high-density genetic map and its application for leaf shape QTL mapping in poplar. Planta 248, 1173–1185. doi: 10.1007/s00425-018-2958-y

PubMed Abstract | CrossRef Full Text | Google Scholar

Xie, Y., Arnold, R. J., Wu, Z., Chen, S., Du, A., Luo, J. (2017). Advances in eucalypt research in China. Front. Agr. Sci. Eng. 4, 380. doi: 10.15302/J-FASE-2017171

CrossRef Full Text | Google Scholar

Xie, C., Mao, X., Huang, J., Ding, Y., Wu, J., Dong, S., et al. (2011). KOBAS 2.0: A web server for annotation and identification of enriched pathways and diseases. Nucleic Acids Res. 39, W316–W322. doi: 10.1093/nar/gkr483

PubMed Abstract | CrossRef Full Text | Google Scholar

Yang, H., Weng, Q., Li, F., Zhou, C., Li, M., Chen, S., et al. (2018a). Genotypic variation and genotype-by-Environment interactions in growth and wood properties in a cloned Eucalyptus urophylla × E. tereticornis family in southern China. For. Sci. 64, 225–232. doi: 10.1093/forsci/fxx011

CrossRef Full Text | Google Scholar

Yang, Y., Xuan, L., Yu, C., Wang, Z., Xu, J., Fan, W., et al. (2018b). High-density genetic map construction and quantitative trait loci identification for growth traits in (Taxodium distichum var. distichum × T. mucronatum) × T. mucronatum. BMC Plant Biol. 18, 263. doi: 10.1186/s12870-018-1493-0

PubMed Abstract | CrossRef Full Text | Google Scholar

Yordanov, Y. S., Regan, S., Busov, V. (2010). Members of the LATERAL ORGAN BOUNDARIES DOMAIN transcription factor family are involved in the regulation of secondary growth in Populus. Plant Cell 22, 3662–3677. doi: 10.1105/tpc.110.078634

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Y., Han, X., Sang, J., He, X., Liu, M., Qiao, G., et al. (2016). Transcriptome analysis of immature xylem in the Chinese fir at different developmental phases. PeerJ 4, e2097. doi: 10.7717/peerj.2097

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Y., Li, J., Li, C., Chen, S., Tang, Q., Xiao, Y., et al. (2022). Gene expression programs during callus development in tissue culture of two Eucalyptus species. BMC Plant Biol. 22, 1. doi: 10.1186/s12870-021-03391-x

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhong, R., Richardson, E. A., Ye, Z.-H. (2007). The MYB46 transcription factor is a direct target of SND1 and regulates secondary wall biosynthesis in Arabidopsis. Plant Cell 19, 2776–2792. doi: 10.1105/tpc.107.053678

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhu, X., He, J., Zhou, C., Weng, Q., Chen, S., Bush, D., et al. (2022). Xylem transcriptome analysis in contrasting wood phenotypes of Eucalyptus urophylla × tereticornis hybrids. Forests 13, 1102. doi: 10.3390/f13071102

CrossRef Full Text | Google Scholar

Zimin, A., Stevens, K. A., Crepeau, M. W., Holtz-Morris, A., Koriabine, M., Marçais, G., et al. (2014). Sequencing and assembly of the 22-Gb loblolly pine genome. Genetics 196, 875–890. doi: 10.1534/genetics.113.159715

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: Eucalyptus urophylla × tereticornis, stable QTL, multi-environment trials, candidate genes, genotyping by sequencing

Citation: Zhu X, Weng Q, Bush D, Zhou C, Zhao H, Wang P and Li F (2023) High-density genetic linkage mapping reveals low stability of QTLs across environments for economic traits in Eucalyptus. Front. Plant Sci. 13:1099705. doi: 10.3389/fpls.2022.1099705

Received: 16 November 2022; Accepted: 28 December 2022;
Published: 18 January 2023.

Edited by:

Hanwei Mei, Shanghai Agrobiological Gene Center, China

Reviewed by:

Qingzhang Du, Beijing Forestry University, China
Rodrigo R. Amadeu, Bayer Crop Science, United States

Copyright © 2023 Zhu, Weng, Bush, Zhou, Zhao, Wang and Li. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Fagen Li, lifagen@caf.ac.cn

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.