Skip to main content

ORIGINAL RESEARCH article

Front. Plant Sci., 18 October 2022
Sec. Plant Breeding

Genome-wide association study for resistance to Pseudomonas syringae pv. garcae in Coffea arabica

Caroline Ariyoshi,Caroline Ariyoshi1,2Gustavo Csar Sant&#x;anaGustavo César Sant’ana3Mariane Silva Felicio,Mariane Silva Felicio2,4Gustavo Hiroshi SeraGustavo Hiroshi Sera2Livia Maria Nogueira,Livia Maria Nogueira1,2Lucas Mateus Rivero RodriguesLucas Mateus Rivero Rodrigues5Rafaelle Vecchia Ferreira,Rafaelle Vecchia Ferreira1,2Bruna Silvestre Rodrigues da Silva,Bruna Silvestre Rodrigues da Silva1,2Mrio Lúcio Vilela de ResendeMário Lúcio Vilela de Resende6Suzete Aparecida Lanza DestfanoSuzete Aparecida Lanza Destéfano7Douglas Silva DominguesDouglas Silva Domingues8Luiz Filipe Protasio Pereira,,*Luiz Filipe Protasio Pereira1,2,9*
  • 1Programa de pós-graduação em Genética e Biologia Molecular, Universidade Estadual de Londrina (UEL), Centro de Ciâncias Biológicas, Londrina, Brazil
  • 2Área de Melhoramento Genético e Propagação Vegetal, Instituto de Desenvolvimento Rural do Paraná (IDR-Paraná), Londrina, Brazil
  • 3Tropical Melhoramento & Genética (TMG), Londrina, Brazil
  • 4Programa de pós-graduação em Ciências Biológicas (Genética), Universidade Estadual Paulista “Júlio de Mesquita Filho“ (UNESP), Instituto de Biociências, Campus de Botucatu, Botucatu, Brazil
  • 5Centro de Café Alcides Carvalho, Instituto Agronômico (IAC), Campinas, Brazil
  • 6Departamento de Fitopatologia, Universidade Federal de Lavras (UFLA), Lavras, Brazil
  • 7Instituto Biológico (IB), Campinas, Brazil
  • 8Departamento de Genética, Escola Superior de Agricultura Luiz de Queiroz, Universidade de São Paulo (USP), Piracicaba, Brazil
  • 9Empresa Brasileira de Pesquisa Agropecuária (EMBRAPA-Café), Brasília, Brazil

Bacteria halo blight (BHB), a coffee plant disease caused by Pseudomonas syringae pv. garcae, has been gaining importance in producing mountain regions and mild temperatures areas as well as in coffee nurseries. Most Coffea arabica cultivars are susceptible to this disease. In contrast, a great source of genetic diversity and resistance to BHB are found in C. arabica Ethiopian accessions. Aiming to identify quantitative trait nucleotides (QTNs) associated with resistance to BHB and the influence of these genomic regions during the domestication of C. arabica, we conducted an analysis of population structure and a Genome-Wide Association Study (GWAS). For this, we used genotyping by sequencing (GBS) and phenotyping for resistance to BHB of a panel with 120 C. arabica Ethiopian accessions from a historical FAO collection, 11 C. arabica cultivars, and the BA-10 genotype. Population structure analysis based on single-nucleotide polymorphisms (SNPs) markers showed that the 132 accessions are divided into 3 clusters: most wild Ethiopian accessions, domesticated Ethiopian accessions, and cultivars. GWAS, using the single-locus model MLM and the multi-locus models mrMLM, FASTmrMLM, FASTmrEMMA, and ISIS EM-BLASSO, identified 11 QTNs associated with resistance to BHB. Among these QTNs, the four with the highest values of association for resistance to BHB are linked to g000 (Chr_0_434_435) and g010741 genes, which are predicted to encode a serine/threonine-kinase protein and a nucleotide binding site leucine-rich repeat (NBS-LRR), respectively. These genes displayed a similar transcriptional downregulation profile in a C. arabica susceptible cultivar and in a C. arabica cultivar with quantitative resistance, when infected with P. syringae pv. garcae. However, peaks of upregulation were observed in a C. arabica cultivar with qualitative resistance, for both genes. Our results provide SNPs that have potential for application in Marker Assisted Selection (MAS) and expand our understanding about the complex genetic control of the resistance to BHB in C. arabica. In addition, the findings contribute to increasing understanding of the C. arabica domestication history.

Introduction

Coffee (Coffea sp.) is one of the most important crops in the world, especially for tropical countries. Although the genus Coffea contains 124 species, only two have economic importance, C. arabica L. and C. canephora Pierre, corresponding to 59.96 and 40.04% of the world production, respectively (Davis et al., 2011; International Coffee Organization, 2020). C. arabica is an autogamous and tetraploid species (2n = 4x = 44) with two diploid ancestors: C. canephora and C. eugenioides (Lashermes et al., 1999; Clarindo and Carvalho, 2008). It was originated in South-West Ethiopia and South Sudan region and traditionally is cultivated at high altitudes. On this conditions, C. arabica is known to provide better beverage quality (Bertrand et al., 2006; Davis et al., 2012; Worku et al., 2018).

Disease incidence is one of the main problems in the cultivation of C. arabica, affecting both production and quality. Among the bacterial diseases, we highlight the bacterial halo blight (BHB) caused by Pseudomonas syringae pv. garcae (Amaral et al., 1956). This disease is prevalent at high altitudes, due to cold winds (Maciel et al., 2018). Worldwide, there are reports of its occurrence in Kenya, Ethiopia, China, and Brazil (Ramos and Shavdia, 1976; Adugna et al., 2012; Xuehui et al., 2013; Rodrigues et al., 2017a). BHB management, based on windbreaks, is not always possible on coffee farms and management based on chemicals is environmentally undesirable. Both solutions increase the production costs and, generally, are not highly efficient in field conditions. Thus, the use of cultivars resistant to BHB is a better alternative for coffee producers.

During the plant-pathogen interaction, once the pathogens overcome the mechanical defense barriers of the plants, resistant plants activate receptors that can recognize the pathogen, activating signaling pathways that drive the defense response gene expression. These defense genes can reduce pathogen infection and colonization through their antimicrobial and cell wall enforcement properties (Andersen et al., 2018). At the first level of recognition, pathogen conserved molecular patterns, known as pathogen-associated molecular patterns (PAMPs), can be detected by host proteins, known as pattern recognition receptors (PRRs), which are capable of activating PAMP-triggered immunity (PTI). The plasma membrane localized plant receptors, such as receptor-like kinases (RLKs) and receptor-like proteins (RLPs) are examples of PRR proteins (Macho and Zipfel, 2015). However, some specialized pathogens produce effectors that can inhibit the PTI (Boller and Felix, 2009).

The second level of recognition involves R proteins capable of detecting effectors and activates effector-triggered immunity (ETI) (Jones and Dangl, 2006; Yuan et al., 2021). Among the R proteins, the class with the greatest representativeness is the nucleotide binding sites leucine-rich repeats (NBSs-LRRs) (Coll et al., 2011). The NBS-LRR and effector interaction can occur directly or indirectly. In the indirect way, the NBS-LRR monitors the binding of the effector with a “decoy” protein (Cui et al., 2015). Both ETI and PTI can activate the hypersensitive response (HR) (Kushalappa et al., 2016). Furthermore, the ETI robustness depends on the PTI (Yuan et al., 2021). The HR is considered as qualitative resistance controlled by few genes that have major effect (R genes) (Jones et al., 2016), and the qualitative resistance is defined by two categories: resistant and susceptible plants (Corwin and Kliebenstein, 2017).

The third line of defense is the absence or weakening of PTI and/or ETI, without HR (Niks et al., 2015). In this type of resistance, some antimicrobial and cell wall enforcement products are unavailable or available in reduced amounts, which allows the advance of pathogens (Boyd et al., 2012). This type of resistance can occur in a constitutive or an induced way and correspond to quantitative resistance (Kushalappa and Gunnaiah, 2013). In the quantitative resistance, the individuals can be divided into a continuous distribution between the resistance and susceptibility, which is related to the number and amount of defense products available (Kushalappa and Gunnaiah, 2013; Corwin and Kliebenstein, 2017). The genes controlling quantitative resistance are located at multiple loci with variable effects (Pilet-Nayel et al., 2017).

Due to its reproductive biology and relatively recent origin, C. arabica has low genetic diversity (Lashermes et al., 2000). In addition, most C. arabica cultivars are derived from only two varieties Typica and Bourbon, a factor responsible for a genetic bottleneck during the cultivar development (Anthony et al., 2001). Consequently, several cultivars are susceptible to diseases and pests (van der Vossen et al., 2015).

In contrast to the lack of genetic diversity in C. arabica cultivars, greater diversity is found in C. arabica Ethiopian accessions. Due to this diversity, a survey organized by the Food and Agriculture Organization (FAO) between 1964 and 1965, collected these accessions in Ethiopia. The collection took place in two regions, the East and West of the Great Rift Valley (FAO: Food and Agriculture Organization, 1968). During the C. arabica domestication process, there was, probably, a movement of plants from the West to the East of the Great Rift Valley (Montagnon and Bouharmont, 1996; Scalabrin et al., 2020).

Several studies, based on this C. arabica Ethiopian germplasm, identified sources of desirable traits, such as resistance to BHB (Mohan et al., 1978); resistance to nematodes Meloidogyne paranaensis, M. incognita, and M. exigua (Anzueto et al., 2001; Fatobene et al., 2017; Holderbaum et al., 2020); resistance to coffee berry disease (van der Vossen and Walyaro, 2009); drought tolerance (Queiroz-Voltan et al., 2014); biochemical compounds which influence beverage quality (Bertrand et al., 2006); low caffeine content (Silvarolla et al., 2004); and other chemical compounds (Scholz et al., 2016).

A large number of genetic recombination historical events present among unrelated individuals allows high-resolution mapping through Genome-Wide Association Studies (GWAS) (Migicovsky and Myles, 2017). Recent association maps, using C. arabica Ethiopian accessions, identified alleles associated with the chemical composition of green grains (chlorogenic acids levels, caffeine, total sugars, sucrose, lipids, proteins, and tannins) (Sant’ana et al., 2018).

In this work, to explore the genetic diversity of C. arabica Ethiopian accessions as well as of selected cultivars, we perform a GWAS for C. arabica and P. syringae pv. garcae interaction. We also identify and validate candidate genes involved in qualitative resistance to BHB by expression analysis and discuss whether the QTNs associated with resistance to BHB may be inserted in genomic regions that were relevant during the C. arabica domestication history.

Materials and methods

Plant materials

A total of 132 C. arabica genotypes were used, including 120 Ethiopian accessions from the FAO collection (FAO: Food and Agriculture Organization, 1968), the BA-10 genotype, and 11 cultivars: IPR 99, IPR 100, IPR 101, IPR 102, IPR 103, IPR 104, IPR 105, IPR 107, Catuaí Vermelho IAC 99, Bourbon Vermelho, and IAPAR 59 (Supplementary Table 1).

Phenotype data, for resistance to BHB, were accessed from two previous works conducted at the Instituto de Desenvolvimento Rural do Paraná (23° 22’ S, 51° 10’ W, 585 m asl) in Londrina, Paraná State, Brazil (Mohan et al., 1978; Ito et al., 2008). The phenotyping of C. arabica Ethiopian accessions, Bourbon Vermelho cultivar, and BA-10 genotype was carried out in the field with two replicates in plants that were approximately 1-year-old. The isolate number 586 of P. syringae pv. garcae was used and evaluations were carried out 17 and 35 days after inoculation (dai). The two evaluations obtained the same results (Mohan et al., 1978). The phenotyping of IPR 99, IPR 100, IPR 101, IPR 102, IPR 103, IPR 104, IPR 105, IPR 107, IAPAR 59, and Catuaí Vermelho IAC 99 cultivars were also evaluated in the field, with nine replications in plants approximately 10-months-old, with natural infections (Ito et al., 2008). In those studies, a scale with a score from 0 to 5 was used, where 0 corresponds to highly resistant plants, 1 to resistant, 2 to moderately resistant, 3 to moderately susceptible, 4 to susceptible, and 5 to highly susceptible (Supplementary Figure 1).

The current work used the GBS data from DNA samples of these 132 C. arabica accessions (Sant’ana et al., 2018). This single-end sequencing was performed using the restriction enzyme PstI, through the platform Illumina HiSeq2000 in partnership with the Genomic Diversity Facility LIMS, at Cornell University (Ithaca, NY - USA). This work used the GBS data (1,083,002 tags) aligned to the C. arabica Et39 reference genome (Salojärvi et al., 2021), performed by Felicio (2020) which allowed identification of 159,000 SNPs.

SNP filtering and linkage disequilibrium analyses

The 159,000 SNPs were filtered using TASSEL software version 5.2.53 (Bradbury et al., 2007). The filtering was performed with the parameters of the minimum allele frequency greater than 0.05 (MAF> 0.05), call rate> 0.8, and removal of monomorphic and indel sites. The data were also imputed by LD kNNi software, which is based on the k-nearest neighbor method (Money et al., 2015). The SNPs obtained after this filter quality were used for the population structure study and for the GWAS.

For the linkage disequilibrium (LD) analysis, the panel of 159,000 SNPs was filtered through the software TASSEL, version 5.2.53. The filters used were: MAF> 0.05, call rate> 0.25, and removal of monomorphic and indel sites. The SNPs corresponding to scaffolds that did not have a defined position during the genome assembly (Chr zero) were also removed. The squared correlation coefficients (r2) were calculated on sliding windows with 50 adjacent SNPs, in TASSEL version 5.2.53, and used to evaluate the extent of LD decay. The value of LD distance (in bp) was evaluated by a non-linear regression, at r2 =0.2, in R studio software (RStudio Team, 2020).

Population structure study

A dissimilarity matrix was built and used for hierarchical analyses, with the complete linkage method and the Euclidean distance, by the R package cluster, and visualized graphically through the factoextra package (Maechler et al., 2019; RStudio Team, 2020; Kassambara and Mundt, 2020).

A Principal Component Analysis (PCA) was performed using TASSEL software, version 5.2.53 (Bradbury et al., 2007). The dispersion plot from the first and second Principal Components (PCs) was conducted using the R package cluster and visualized graphically using the package ggfortify (Horikoshi and Tang, 2016).

Population structure was also analyzed using the Bayesian method, implemented in STRUCTURE 2.3.4 (Pritchard et al., 2000). We predefined the number of genetic clusters K from 2 to 10 in the population. The program was set on 10,000 as burn-in iteration, followed by 10,000 Markov chain Monte Carlo (MCMC) replications with 20 runs per K. We used the ΔK criterion in Structure Harvester software to estimate the upper-most level of structure (Evanno et al., 2005; Earl and von Holdt, 2012).

Gwas

The associative mappings were conducted using a single-locus model and four multi-locus models. The single-locus model used was the mixed linear model (MLM), through the software TASSEL 5.2.53 (Yu et al., 2006; Bradbury et al., 2007). For MLM, the association threshold established, to reduce false positive associations, was Bonferroni. For the multi-locus approach, we used four models within the R package mrMLM: “random-SNP-effect MLM (rMLM) and a multi-locus mrMLM” (mrMLM) (Wang et al., 2016), FASTmrMLM (Tamba and Zhang, 2018), “fast multi-locus random-SNP-effect EMMA” (FASTmrEMMA) (Wen et al., 2017), and “integrative sure independence screening EM-Bayesian LASSO” (ISIS EM-BLASSO) (Tamba et al., 2017).

These multi-locus models have two stages. In the first stage, a single-locus model and a modified Bonferroni association significance threshold are used, producing an intermediate result. In the second stage, for the multi-locus model implementation, it is necessary to reduce the data dimensionality. Thus, only the SNPs with the greatest potential for association selected in the first stage composed the data for the second stage. The critical values for SNPs to integrate data for the second stage were: p-value ≤ 0.01, 0.01, 0.005, and 0.01 for mrMLM, FASTmrMLM, FASTmrEMMA, and ISIS EM-BLASSO, respectively. In the second stage, the association threshold for all models was the LOD score = 3. The QTNs associated above LOD = 3 corresponded to the final result.

The single-locus model was adjusted using the first five PCs based on the dissimilarity matrix and kinship matrix (K matrix). To verify the best adjusted performance for the multi-locus models, these were corrected by K matrix, K matrix + the first five PCs based on the 11,290 SNPs and K matrix + Q matrix. The Q matrix was based on the highest level of population structure (K = 2) identified from software STRUCTURE. For the single-locus MLM model, the K matrix was calculated using TASSEL 5.2.53 software, and for the multi-locus models, using the mrMLM package.

From the HapMap genotyping file, the presence or absence of the QTNs identified in the GWAS was performed in the 132 C. arabica accessions, with its respective resistance to BHB phenotype score. With this identification, the coefficient of correlation between the QTNs associated with resistance to BHB was calculated. The coefficient of correlation between each QTN identified in GWAS and the resistance to BHB, in the 132 C. arabica accessions, was also calculated. The coefficient of correlation calculation and the graphical visualization were made using the R packages Hmisc and corrplot, respectively (Harrell and Dupont, 2020; Wei and Simko, 2021).

Identification of candidate genes

To understand the molecular basis of the C. arabica defense to BHB, we investigated the genomic regions of associated QTNs for possible candidate genes. A search for candidate genes was performed using the C. arabica Et039 genome functional annotation (Salojärvi et al., 2021), according to the LD threshold, upstream and downstream to the associated QTNs positions. Functional annotation and genomic position of the C. arabica genes was retrieved and examined from the C. arabica Et039 genome functional annotation. Subsequently, proteins with biotic stress response function were literature mined from annotated information.

Plant materials for RT-qPCR

Plants of 5-months-old of C. arabica cultivars Catuaí Vermelho IAC 99 (susceptible), IAPAR 59 (quantitative resistance), and IPR 102 (qualitative resistance) were submitted to infection with P. syringae pv. garcae through the method established by Rodrigues et al. (2017b). The inoculum of P. syringae pv. garcae grown on agar for 48h (28 °C), was diluted in saline (0.85% NaCl) and standardized by spectrophotometry (A600 = 0.25) containing 108 UFC/ml. Two pairs of the first expanded leaves of seedlings were inoculated with punctures through needles, previously dipped in the inoculum, at two points on each side of the central rib.

Leaves were collected before the inoculation at 0 hours (h) and, 10 minutes (min), 30 min, 1 h, 3 h, and 6 h, time after inoculation (tai), for RNA extraction. The experimental design for RT-qPCR was conducted with 3 plants of each cultivar for each collection time.

Two inoculated leaves, in each plant, were preserved to verify the inoculum efficiency and the disease evaluation was performed at 21 (dai). The damage in the inoculated plants was estimated using a scale from 0 to 5, where 0 = resistant, absence of symptoms; 1 = moderately resistant, initial water-soaking, up-to 10% of leaf area inoculated with symptoms; 2 = susceptible, water-soaking up to approximately 25% of affected area; 3 = up to 50% of the area inoculated with symptoms; 4 = up to 75% of the affected inoculated area; 5 = more than 75% of the inoculated area necrotic from the disease (Rodrigues et al., 2017b).

Expression profile of candidate genes by RT-qPCR

The RNA extraction was carried out following the protocol of Chang et al. (1993). The cDNAs were synthesized from 2.5 μg of total RNA using the reverse transcriptase enzyme from the SuperScript® III First-Strand Synthesis SuperMix kit (Invitrogen), according to the manufacturer’s instructions, with a final volume of 20 μl.

Specific pairs of primers were designed for the cDNA corresponding to g000 (Chr_0_434_435) and g010741 genes using Primer Express software (Thermo Fisher Scientific) (Supplementary Table 5). Primer efficiencies were performed through the standard curve of cDNA dilutions: 1:2, 1:4, 1:8, 1:16, 1:32, and 1:64. Every measurement of dilutions was performed in triplicate. The corresponding qPCR efficiencies (E) were calculated for primer pairs with the 7500 Fast Real-Time PCR System (Applied Biosystems), according to the equation E = (10-1/slope - 1) × 100 (Rutledge and Stewart, 2008).

The transcriptional profile was evaluated using qPCR (7500 Fast Real-Time PCR System, Applied Biosystems). The reaction volumes was 25 μL containing 12.5 μL of 2x SYBR® Green/ROX qPCR Master Mix (Applied Biosystems), 0.5 μL of each primer (10 μM), 10.5 μL of treated DEPC water, and 1 μL of cDNA diluted to a concentration of 20 ng/μL.

The reactions were prepared in technical triplicates and the thermocycling parameters used were: 10 min at 95 °C, followed by 40 cycles of amplification of 95 °C for 30 seconds and 60 °C for 60 seconds. Dissociation curves were analyzed to verify the specificity of the primers in the amplification reaction.

The relative gene transcription was assessed using ΔΔCt = ΔCt (sample) - ΔCt (normalizer), where the GAPDH gene was used as housekeeping to normalize the data (Barsalobres-Cavallari et al., 2009). The calibrator used to compare the transcriptional activities pattern, between the times after inoculation, was the 0 h for each cultivar. The transcript data were subjected to statistical analysis of variance (ANOVA) and the Tukey test, both at the level of 5% significance, through R studio software (RStudio Team, 2020).

Haplotype analysis

From the genomic region corresponding to the g010741 gene (CC-NBS-LRR) of C. arabica Et039 and the respective homologous genomic regions of C. arabica Caturra and, C. arabica var. Geisha, a haplotype analysis was performed. For this, the information available in the genome databases of C. arabica Et039 (Salojärvi et al., 2021), C. arabica Caturra (NCBI BioProject PRJNA506972), and C. arabica var. Geisha (Phytozome genome ID 453), was used.

The genomic region corresponding to the C. arabica Et039 g010741 gene (Chr2_sg_E, position 32,109,397 bp – 32,111,166 bp) was used to identify the homologous genomic regions in Caturra and Geisha, through the BLASTn tool. Multiple alignment between the nucleotide sequences of the C. arabica Et039 g010741 gene and the homologous nucleotide sequences of Caturra, and Geisha was carried out to identify SNPs, using MUSCLE (Edgar, 2004). The respective protein sequences were also aligned to identify non-synonymous SNPs (nsSNPs). SNPs inserted into conserved domains were investigated. For this, a protein sequence of 1,280 aa corresponding to the g010741 gene was retrieve from the C. arabica Et039 functional annotation and used to the identification of conserved domains and their respective genomic positions, through the InterPro (Blum et al., 2021).

Near to the g010741 gene are the QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G associated with resistance to BHB. Thus, in addition to the haplotype analysis described above, the identification of standard alleles or alternative alleles for the QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G, in the three C. arabica genotypes, was made. From the QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G positions in the Et039 genome (reference genome used to align the GBS data in this work), the identification of standard alleles or alternative alleles was made. Using the nucleotide sequence (60 bp) surrounding this QTNs in Et039, a search for homology in Caturra and Geisha genotypes, was performed. The search for homology was performed using the BLASTn tool.

Relationship between the loci involved in resistance to BHB and the C. arabica domestication process

The frequency of QTNs identified in the GWAS was determined in the clusters “wild”, “domesticated”, and “cultivar” identified in the Bayesian approach population structure at K = 3. For this, C. arabica Ethiopian accessions classified as admixed were disregarded. The frequency of genotypes resistant to the BHB, based in a scale with a score from 0 to 5, was also determined in the clusters

Results

SNP characteristics

After the quality filters and imputation, we selected 11,290 SNPs for population structure analysis and GWAS. From this total, 5,024 are distributed in the C. canephora subgenome, 5,011 in the C. eugenioides subgenome, and 1,255 in chromosome (Chr) zero. The Chr zero corresponds to scaffolds that did not have a defined position during the genome assembly. The SNPs/Megabase (Mb) frequency was 13.57 and 11.63 for the C. canephora and C. eugenioides subgenome, respectively. Among the chromosomes a large difference was observed in the SNP/Mb frequency. The Chr2 of the C. canephora subgenome showed the highest density of SNPs (17.81) and the Chr9 of the C. eugenioides subgenome the lowest (8.41) (Supplementary Figure 2 and Supplementary Table 2).

Population structure

The population used in this work is compose of 111 C. arabica Ethiopian accessions collected at the West of the Great Rift Valley, nine collected at the East of the Great Rift Valley, 11 C. arabica cultivars, and the BA-10 genotype.

The clustering based on the dissimilarity matrix of genetic distance of the 132 C. arabica accessions (Supplementary Figure 4), through the hierarchical approach, defined 3 clusters (Figure 1A and Supplementary Table 1). The cluster in blue is composed of 32 accessions, most of them collected in forests, which represents a lower domestication level, and the BA-10 genotype, which has a C. liberica introgression. The cluster in red is composed of 88 accessions collected, predominantly, in places with the description of naturalized, domesticated, farm open field and shadow. Naturalized, domesticated, and, mainly, farm open field represents a high domestication level. The cluster in green is composed of the C. arabica cultivars. Thus, the blue cluster was named “wild”, the red cluster “domesticated”, and the green cluster “cultivar”. Among the nine accessions collected at the East of the Great Rift Valley, one is in the “wild” cluster and eight are in the “domesticated” cluster.

FIGURE 1
www.frontiersin.org

Figure 1 Genetic population analysis among 132 Coffea arabica genotypes using 11,290 SNPs. (A) Dendrogram; (B) principal coordinate analysis (PCA). Each point represents an accession, n = 132. In the PCA the accessions were colored according to the cluster: red “domesticated”, blue “wild”, and green “cultivar”. (C) Bar plots of the estimated membership coefficient of the 132 C. arabica accessions for K = 2, K = 3, K = 4 and, K = 5, using STRUCTURE.

The PCA scatter plot also showed three clusters, with accessions grouped according to the domestication level. The “wild” cluster in blue is composed of 33 accessions, predominantly, collected in forests, and the BA-10 genotype. The “domesticated” cluster in red is composed of 85 accessions, which were predominantly characterized as domesticated. However, different from the hierarchical analysis result, the “cultivar” cluster in green is composed of cultivars and more two C. arabica Ethiopian accessions collected in “farm open field”. Among the accessions collected at the East side of the Great Rift Valley, one is in the “wild” cluster, seven are in the “domesticated” cluster, and one is in the “cultivar” cluster (Figure 1B and Supplementary Table 1).

According to the Bayesian approach, using STRUCTURE software and Evanno’s criterion, the upper level of the population subdivision were K = 2. In addition, smaller peaks were observed at K = 3, and at K = 5 (Supplementary Figure 5), which might indicate another informative population structure. Therefore, the STRUCTURE results at K = 2, K =3, K = 4, and K = 5 were subject population genetics analyses. The population genetics analyses were based on the membership coefficient (≥ 0.6), so that the accessions could be assigned to a specific group. The membership coefficients of the C. arabica cultivars and C. arabica Ethiopian accessions at K =2, K = 3, K = 4, and K = 5 are available in the Supplementary Material 1.

At K = 2, the 11 cultivars and 30 C. arabica Ethiopian accessions, most of them characterized as domesticated, are grouped in the same cluster (in red). The second cluster (in green) is composed of the 32 most wild C. arabica Ethiopian accessions and 24 domesticated accessions. For this sub-populations structure, 34 C. arabica Ethiopian accessions and the BA-10 genotype (26,51%) were categorized as admixed. Among the accessions collected at the East of the Great Rift Valley, seven are in the red cluster, and two were classified as admixed (Figure 1C and Supplementary Figure 6).

A clearer separation of the domesticated C. arabica accessions from the cultivars, and of the domesticated C. arabica accessions from the most wild accessions could be observed at K = 3. The green cluster is composed of 11 cultivars and 5 domesticated accessions. The cluster in red is composed of 19 domesticated accessions and the BA-10 genotype. The blue cluster is composed of 33 accessions, predominantly, collected in forest. Thus, the green cluster corresponds to the “cultivar” cluster, the red to the “domesticated” cluster, and the blue to the “wild” cluster. At this population structure level, 63 (47,72%) accessions were categorized as admixed. Among the accessions collected at the East of the Great Rift Valley, one is in the “domesticated” cluster, three in the “cultivar” cluster, and five were classified as admixed (Figure 1C and Supplementary Figure 6). Although five accessions collected in the East were classified as admixed, it is possible to observe in three of these accessions a membership coefficient > 0.5 of the “cultivar” cluster. This result reinforces the genetic proximity between the cultivars and the accessions collected at the East of the Great Rift Valley.

The clustering pattern obtained at K = 4 revealed a “wild” cluster in blue with the BA-10 genotype and 28 accessions, predominantly, collected in forest, and a “cultivar” cluster in green with 11 cultivar and three domesticated accessions. However, at K = 4, were identified two “domesticated” clusters in red and in yellow, with 22 and 8 domesticated accessions, respectively. We did not identify any characteristic that could differentiate the accessions grouped in different “domesticated” clusters. For this four sub-populations, 59 (44,69%) accessions were classified as admixed. Among the accessions collected at the East of the Great Rift Valley, one is in the “cultivar” cluster, and eight were classified as admixed (Figure 1C and Supplementary Figure 6). In the same way as was observed at K = 3, three accessions collected at the East and classified as admixed have a membership coefficient > 0.5 of the “cultivar” cluster.

At K = 5, a “wild” cluster is represented in purple with 26 accessions, predominantly, collected in forest and the BA-10 genotype, and a “cultivar” cluster is represented in green with 11 cultivars and two domesticated accessions. Three “domesticated” clusters, in red, blue and yellow, with 11, 8, and 8 accessions, respectively, were identified. Here, we also did not identify characteristics that could distinguish the accessions grouped in the three different “domesticated” clusters. At K = 5, 63 (49,24%) accessions were classified as admixed. Among the accessions collected at the East of the Great Rift Valley, one is in the “domesticated” cluster in blue, and eight were classified as admixed (Figure 1C and Supplementary Figure 6). For this population structure level, only one accession collected in the East of the Great Rift Valley and classified as admixed has a membership coefficient > 0.5 of the “cultivar” cluster.

Although the higher delta K value was identified at K = 2, the population structure identified at K = 3 was suitable for grouping the C. arabica accessions according to their domestication level characteristics. This is in agreement with the hierarchical analysis and PCA results. In addition, the population structure level identified by the Bayesian approach at K =3, was suitable to show the most genetic similarity between accessions collected at the East of the Great Rift Valley and the cultivars. A large difference was observed in the number of accessions that compose the “domesticated” cluster, between the PCA analysis and the Bayesian approach. This is mainly because many accessions were classified as admixed in the Bayesian approach.

Gwas

From 11,290 SNPs obtained after quality filters, 11 QTNs associated with the C. arabica response to BHB, through the single-locus MLM model and the four multi-locus models, mrMLM, FASTmrMLM, FASTmrEMMA, and ISIS EM-BLASSO, were identified (Table 1 and Figures 2 and 3A). The single-locus MLM model identified 4 QTNs. For the multi-locus GWAS approaches, a better fit of the data on Q-Q plot was observed in the model corrected with K matrix + PCA (Figure 3B). The multi-locus models mrMLM, FASTmrMLM, and ISIS EM-BLASSO, adjusted with K matrix + PCA, identified 6, 4, and 5 QTNs, respectively. The FASTmrEMMA did not identify associated QTNs.

TABLE 1
www.frontiersin.org

Table 1 Description of QTNs associated with resistance to BHB in C. arabica through the MLM, mrMLM, FASTmrMLM, and ISIS EM-BLASSO models.

FIGURE 2
www.frontiersin.org

Figure 2 Manhattan plot and Q-Q plot corresponding to single-locus GWAS MLM for C. arabica and P. syringae pv. garcae interaction. The manhattan plot indicate the -log10(p-value) of SNP across the genome (y-axis) plotted against their respective position on each chromosome (x-axis). The blue line represents the Bonferroni corrected threshold. The chromosome (Chr) zero correspond to scaffolds that did not have a defined position during the genome assembly, Chr 1 to 11 correspond to canephora subgenome, Chr 12 to 22 correspond to eugenioides subgenome.

FIGURE 3
www.frontiersin.org

Figure 3 Multi-locus GWAS. (A) Manhattan plots corresponding to mrMLM, FASTmrMLM, and ISIS EM-BLASSO for C. arabica and P. syringae pv. garcae interaction. The manhattan plots indicate the -log10 (p-value) and the LOD score of SNP across the genome (y-axis) plotted against their respective position on each chromosome (x-axis). The chromosome (Chr) zero correspond to scaffolds that did not have a defined position during the genome assembly, Chr 1 to 11 correspond canephora subgenome, Chr 12 to 22 correspond to eugenioides subgenome. (B) Q-Q plots from multi-locus models adjusted by different population structure approaches. Kinship matrix (K matrix), principal component analysis (PCA), and the population structure K = 2 identified from software STRUCTURE (Q matrix).

The QTNs Chr_2_sg_E_32049720_G, Chr_2_sg_E_32049728_G, Chr_0_434_435_15461_A, and Chr_0_434_435_15529_A have the highest association -log10 (p-value) and a high correlation with resistance to BHB (Table 1, Figure 4). While the other QTNs, identified only through the multi-locus models: Chr_1_sg_C_27081268_C, Chr_5_sg_C_29867225_G, Chr_7_sg_C_1131696_C, Chr_10_sg_C_439213_G, Chr_11_sg_C_11719181_C, Chr_5_sg_E_19112445_C, and Chr_7_sg_E_13418072_C showed a lower correlation to the level of resistance to BHB (Table 1, Figure 4).

FIGURE 4
www.frontiersin.org

Figure 4 Correlation between the 11 QTNs identified in the GWAS and the resistance to BHB phenotype. The closer to zero, the lower the correlation, positive (blue) or negative (red), which depends on the associated QTN effect.

Furthermore, only the IPR 102 cultivar and some Ethiopian accessions like E287 with high resistance to BHB have the QTNs: Chr_2_sg_E_32049720_G, Chr_2_sg_E_32049728_G, Chr_0_434_435_15461_A, and Chr_0_434_435_15529_A (Supplementary Table 3).

Identification of candidate genes

From the C. arabica Et039 functional annotation, a total of 31 C. arabica genes were identified near to 11 QTNs associated with resistance to BHB (Supplementary Table 4). To consider the gene linked to the QTN, distances approximately based on the LD decay result, were used. The LD decay at r2 = 0.2 was 66,291 bp (Supplementary Figure 3). No genes were identified near to the QTNs Chr_11_sg_C_11719181_C and Chr_5_sg_E_19112445_C. Among the C. arabica genes identified near to the QTNs associated with resistance to BHB, we found eight genes with functional annotation related to response to biotic stress described in the literature (Table 2).

TABLE 2
www.frontiersin.org

Table 2 Functional annotation of genes near to QTNs associated with resistance to BHB in C. arabica identified in this study.

Interestingly, near to the QTN with the highest -log10(p-value) of association, Chr_2_sg_E_32049720_G, we found the g010741 gene with functional annotation for a disease resistance RPP13. This protein is a NBS-LRR, which acts on specific recognition of pathogen and hypersensitivity response (HR). The QTN Chr_0_435_15529_A, also with high -log10(p-value) of association, is linked to a serine threonine- kinase D6PKL2. Another protein of this class, a serine threonine- kinase endoribonuclease IRE1a-like, was identified near to the QTN Chr_10_sg_C_439213_G.

Regarding defense compounds synthesized by plants, we identified C. arabica genes with functional annotation for hydrolase (g009091), glycosidase (g001147), RNA-binding protein linked to callose deposition (g009946), and enzyme involved in alkaloid biosynthesis (g009954). It is worth mentioning that the QTN Chr_1_sg_C_27081268_C is near to g090102, a gene with functional annotation for the Arabinogalactan protein family.

The detailed descriptions of the genes roles in the plants defense mechanism against biotic stresses, as well as their respective literary references, are presented in discussion topic.

Transcriptional profile of candidate genes by RT-qPCR

The genes g000 (Chr_0_434_435) and g010741 (Chr_2_sg_E), predicted for encoding serine/threonine-kinase protein and CC-NBS-LRR, respectively, showed no difference in transcription levels in the absence of the pathogen for the Catuaí Vermelho IAC 99, IAPAR 59, and IPR 102 cultivars (Supplementary Figures 8A, B). However, after inoculation with the pathogen, different transcription levels were observed (Figures 5A, B).

FIGURE 5
www.frontiersin.org

Figure 5 (A) Transcription levels of genes related to BHB resistance in leaves of C. arabica Catuaí Vermelho IAC 99, IAPAR 59, and IPR 102 inoculated with P. syringae pv. garcae. (A) Serine/threonine- kinase g000 (Chr_0_434_435) and (B) CC-NBS-LRR g010741 (Chr_2_sg_E). Data represent means ± standard deviation. Upper case letters compare columns of different genotypes at the same tai and lower-case letters compare columns of the same genotype at different tai. Same letters show no difference by the Tukey test (for P ≤ 0.05). (C) Phenotypic analysis of IPR 102, IAPAR 59, and Catuaí Vermelho IAC 99, 21 days after inoculation.

The three cultivars showed repression of serine/threonine-kinase transcripts from 10 min (tai). However, the repression in Catuaí Vermelho IAC 99 and IAPAR 59 was higher than observed in IPR 102 during the time course of infection. The cultivar Catuaí Vermelho IAC 99 showed a progressive decrease in repression of serine/threonine-kinase transcripts from 30 min to 6h (tai). The cultivar IAPAR 59 also showed a repression decrease at 30 min (tai) but remained stable until 6 h (tai). The serine/threonine-kinase transcript repression in IPR 102 remained stable until 6 h (tai) when it showed an upregulation.

Regarding the CC-NBS-LRR gene, the highly resistant cultivar IPR 102 showed no change in the level of transcription at 10 min (tai) compared to that observed in the absence of the pathogen, but at 30 min (tai) an upregulation peak approximately 30 times greater than that found in the cultivar in the absence of the pathogen was observed. However, at 1 h and 3 h (tai) the IPR 102 returned to show the same level of transcription that was found in the absence of the pathogen for the CC-NBS-LRR gene. The IPR 102 showed another peak of upregulation at 6 h (tai), approximately 50 times higher than the level found in the absence of the pathogen and with statistical difference from the first peak.

The cultivar Catuaí Vermelho IAC 99 showed a pattern of progressive downregulation between 10 min (tai) and 6 h (tai) for CC-NBS-LRR transcripts. The cultivar IAPAR 59 also showed a pattern of downregulation of this gene from 10 min (tai), but similar to the serine/threonine-kinase the level of repression remained stable until 6 h (tai).

For the RT-qPCR analyses, the choice of collection times is based on the results of Cheval et al. (2013) and Nobori et al. (2020). According to these authors, the interaction of Arabidopsis thaliana vs. P. syringae in the early stages, before 6 h (tai), are essential to determine the plant resistance or susceptibility.

All seedlings of Catuaí Vermelho IAC 99, IAPAR 59, and IPR 102 used for qPCR analyses, showed leaves with the score 4 (susceptible), 2 (moderately resistant), and 0 (highly resistant) at 21 (dai), respectively (Figure 5C).

Haplotype analyses

The C. arabica Et039 g010741 gene showed homology with the genomic region 31,924,108 bp – 31,927,645 bp (Chr2_sg_E) of Caturra, and with the genomic region 1,417,586 bp – 1,421,123 bp (Scaffold_627 | HRSCAF=12830) of Geisha genotype. The homologous genomic region identified in Caturra, corresponds to the locus LOC113732450, and in Geisha to the locus Os08g43010. Both LOC113732450 and Os08g43010 have functional annotation for disease resistance RPP13-like protein 1. From the multiple alignment between g010741 (Et039), LOC113732450 (Caturra) and Os08g43010 (Geisha), it was possible to identify 143 SNPs in a sequence region of 3,412 bp. Caturra and Geisha showed the same alleles for these 143 SNPs. The position of each SNP identified in the haplotype analysis is available in Supplementary Material 2. Thus, the alleles identified in Et039 compose the haplotype 1 (Hap1) and the alleles identified in Caturra and Geisha compose the haplotype 2 (Hap 2) (Figure 6B). Among the 143 SNPs identified, 105 correspond to nsSNPs (Figure 6B). The amino acid change corresponding to each nsSNPs is available in Supplementary Material 3.

FIGURE 6
www.frontiersin.org

Figure 6 Haplotype analysis. (A) The C. arabica Et039 g010741 gene (Chr2_sg_E) and their conserved domains with respective positions in pb. (B) The 143 SNPs identified between haplotype 1 (Hap1 – Et039) and haplotype 2 (Hap2 – Caturra and Geisha); and the alleles that compose the haplotypes. The SNPs identified between Hap1 and Hap2, inserted in conserved domains, are highlighted. The position of each SNP identified in the haplotype analysis, in the C. arabica Et039 chromosome 2 eugenioides subgenome, is available in Supplementary Material 2. The 105 nsSNPs are identified with *. The amino acid change corresponding to each nsSNPs is available in Supplementary Material 3. (C) The alleles identified in the QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G in C. arabica Et039 (Hap1) and in Caturra e Geisha (Hap2).

From the protein sequence predicted by the gene g010741, it was possible to identify the conserved domains N-terminal coiled-coil (ID IPR041118), NB-ARC (ID IPR002182), and LRR (ID IPR032675). The positions, in bp, of the conserved domains identified for this gene in the C. arabica Et039 chromosome 2 eugenioides subgenome, was represented in Figure 6A. Among the 143 SNPs that compose the haplotypes, 13 are inserted in the CC domain, 37 in the NB-ARC domain and 71 in the LRR domain (Figure 6B). Among the 105 nsSNPs, 7 are inserted in the CC domain, 24 in the NB-ARC domain and 60 in the LRR domain (Figure 6B).

C. arabica Et039, which represents Hap1, showed the alternative allele G both for the QTN Chr_2_sg_E_32049720_G and for the QTN Chr_2_sg_E_32049728_G. Caturra and Gueisha, which represents Hap2, showed the standard allele C for the QTN Chr_2_sg_E_32049720_G and the standard allele A for the QTN Chr_2_sg_E_32049728_G (Figure 6C). In the Caturra genome the QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G were located in Chr2_sg_E at the positions 31,851,359 bp and 31,851,367 bp, respectively. In the Geisha genome the QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G were located in Scaffold_627|HRSCAF=12830 at the positions 1,493,868 bp and 1,493,859 bp, respectively.

Discussion

Gwas

In this study, we identify 11 QTNs associated with resistance to BHB in C. arabica. Using a multi-locus method, in relation to the single-locus, we identified a greater number of QTNs associated with response to BHB, with a small effect. The advantages of the multi-locus method over single-locus has already been described in other plants and animal species, such as cotton, maize, and pigs (Li et al., 2018; Su et al., 2018; Xu et al., 2018; Ding et al., 2019). This is possible due to the nature of the multi-locus method, which makes a genome multidimensional scan and a simultaneous estimate of the effect of all SNPs (Cui et al., 2018). The ability to identify a greater number of QTNs and QTNs with small effect associated with the trait of interest increase our understanding of the genetic control of complex traits and the estimation of heritability (Ding et al., 2019; Zhang et al., 2019).

Some multi-locus methods were able to overlap the genomic regions associated with resistance to BHB identified through the single-locus method. The QTN Chr_0_435_15529_A was detected using the single-locus MLM and through the multi-locus mrMLM. A complementarity between the results of the single-locus and multi-locus methods was also found for the QTN Chr_2_sg_E_32049720_G, which was identified through the single-locus MLM and the multi-locus mrMLM, FASTmrMLM, and, ISIS EM-BLASSO. Other authors have observed the same ability of multi-locus methods to overlap part of the results obtained by single-locus methods (Misra et al., 2017; Li et al., 2018).

It is worth mentioning that the multi-locus method considers the LD and decreases the number of associated QTNs in the same haplotype (Ding et al., 2019). Probably because of this, the QTN Chr_2_sg_E_32049728_G identified by MLM was not identified in the mrMLM, FASTmrMLM, and ISIS EM-BLASSO, since it is very close to the QTN Chr_2_sg_E_32049720_G. The same was observed for the QTN Chr_0_435_15461_A, in relation to the QTN Chr_0_435_15529_A, in the result of the mrMLM.

Some QTNs were identified through different multi-locus methods, such as the QTN Chr_2_sg_E_32049720_G which was identified by the mrMLM, FASTmrMLM, and ISIS EM-BLASSO, the QTNs Chr_11_sg_C_11719181_C and Chr_5_sg_E_19112445_C identified by the mrMLM, and FASTmrMLM, the QTN Chr_7_sg_C_1131696_C identified by the mrMLM, and ISIS EM-BLASSO, and the QTN Chr_7_sg_E_13418072_C identified by the FASTmrMLM, and ISIS EM-BLASSO, thus, demonstrating the robustness of the association methods as well the reliability of the QTNs.

Identification of candidate genes related to plant defense to BHB

Plant diseases threaten plant production and, consequently, world food security. Identification and understanding of the mechanisms and genetic bases of resistance is essential to achieve greater and more sustainable production. Genes involved in resistance, once identified and validated, can be used to improve the crop based on molecular assisted breeding or using genome editing tools.

Plant resistance is based on metabolites and proteins that directly suppress pathogen development, leading to reduced susceptibility. These metabolites and proteins can be constitutively present or synthesized in plants following pathogen perception (Kushalappa et al., 2016). After the plant proteins recognize the pathogen a hierarchy of downstream genes such as phytohormones (salicylic acid, jasmonic acid, and ethylene), mitogen-activated protein kinases (MAPKs), and transcription factors (TFs), which regulate downstream genes that produce defense metabolites (phytoanticipins, phytoalexins, and complex conjugates that are deposited to reinforce cell walls), and proteins (pathogenesis-related proteins, detoxify toxins proteins, and enforcing cell wall proteins) (Kushalappa and Gunnaiah, 2013).

Some C. arabica genes, near to the QTNs associated with resistance to BHB, identified in this work, have functional annotation for proteins that play roles in the plants defense mechanism against pathogens.

The gene g010741(Chr_2_sg_E) is ortholog to the RPP13, a downy mildew resistance protein (NBS-LRR) originally identified in the Niederzenz (Nd-1) ecotype of Arabidopsis (Bittner-Eddy et al., 2000). Furthermore, Lewis et al. (2013) identified that RPP13L4/ZAR1 is dependent on a nonfunctional kinase, which functions as a “decoy”, for pathogen recognition.

Other genes that act as defense signal receptor and transmitter proteins were also identified. The gene g000 (Chr0) has a functional annotation for serine/threonine-protein kinase D6PKL2. In Zhang et al. (2021) this protein was described as specifically expressed after Fusarium oxysporum infection in Vernicia montana resistant trees. Furthermore, transgenic analysis in Arabidopsis and tomato revealed that D6PKL2 significantly enhanced resistance in both species, whereas the d6pkl2 mutant displayed reduced resistance against F. oxysporum (Zhang et al., 2021). The gene g000156 (Chr_10_sg_C) has orthology with an Arabidopsis gene that encondes a serine/threonine-kinase endoribonuclease IRE1a-like protein. In Arabidopsis, this gene was transcriptionally induced during bacterial pathogen P. syringae pv. maculicola infection and plays a predominant role in the secretion of pathogenesis-related proteins during salicylic acid treatment. Arabidopsis mutant plants without this gene showed enhanced susceptibility to a bacterial pathogen and were deficient in establishing systemic acquired resistance (SAR) (Moreno et al., 2012).

A protein with hydrolytic function, the probable carbohydrate esterase At4g34215, was identified in the g009091 (Chr_1_sg_C) gene functional annotation. Balmant et al. (2015) reported that a carbohydrate esterase exhibited a significant fold change in redox modification and a significant fold change in transcriptional level, during the interaction between tomato resistant genotype (PtoR) and P. syringae pv tomato. The gene g001147 (Chr10_C) also showed functional annotation for a protein with hydrolytic function, a glucan endo-1|3-beta-glucosidase. This enzyme are grouped in the PR-2 family of pathogenesis-related proteins and catalyze β-1,3-glucans cell wall structural component of fungal and bacteria (Bachman and McClay, 1996; Balasubramanian et al., 2012).

Regarding secondary metabolites of plant defense, the functional annotation of the g009954 (Chr_5_sg_C) gene corresponds to an S-norcoclaurine synthase-like. S-norcoclaurine is the ultimate precursor to benzylisoquinoline, an alkaloids that plays a role in the plants defense against herbivores and pathogens (Liscombe et al., 2005; Yogendra et al., 2017).

The g009893 (Chr_5_sg_C) has functional annotation for a UBP1-associated protein 2C. This protein is a RNA-binding and provides hints to another level of regulation posttranscriptionally. In Bove et al. (2008) is reported that transcripts levels of UBA2 increased following mechanical wounding. In another work, using transient expression, Kim et al. (2008) identified that Arabidopsis UBA2 proteins in Nicotiana benthamiana leaves induces a programmed-cell-death/senescing phenotype, while constitutive expression causes an early lethality phenotype in Arabidopsis.

The gene g090102 (Chr1_C) has functional annotation for Arabinogalatans proteins family. Arabinogalactans proteins (AGPs) are structural polysaccharides associated with proteins of the plant cell wall. Plants can secrete and accumulate AGPs at infection sites, creating cross-links with pectin. Pathogens can secret hydrolytic enzymes that damage AGPs. In this way, the AGPs degradation, by the pathogen’s hydrolytic enzymes action, can elicit plant defense response through damage-associated molecular patterns (DAMPs) (Villa-Rivera et al., 2021).

Identification of genomic regions probably involved in qualitative resistance

Major genes that recognize the pathogen and trigger PTI and ETI, resulting in HR, are essential to promote qualitative resistance (Jones and Dangl, 2006; Yuan et al., 2021). In the absence of these genes, other genes with minor effects can promote partial resistance to the pathogen and a decrease in symptom severity (St. Clair, 2010). This partial resistance to the pathogen corresponds to quantitative resistance and produces a continuous distribution of phenotypes between the susceptible and the resistant plant (Corwin and Kliebenstein, 2017).

The elucidation of which genes act on qualitative, quantitative, or both types of resistance is important, since the combination of qualitative and quantitative resistance, in the same genotype, is considered a promising strategy in breeding programs (Pilet-Nayel et al., 2017). This is because quantitative resistance can exercise different selection pressures on the pathogen (Quenouille et al., 2014). In addition, it can reduce the effective population size of the pathogen, which increases the genetic drift and delays the emergence of new virulent forms (Brun et al., 2010). Thus, the effectiveness of resistance conferred by major genes is preserved.

The QTNs Chr_2_sg_E_32049720_G, Chr_2_sg_E_32049728_G, Chr_0_434_435_15461_A, and Chr_0_434_435_15529_A can be associated with qualitative resistance to BHB. Four of our results reinforces this hypothesis. Firstly, the QTNs showed a strong correlation with the resistance phenotype (Figure 4). Secondly, only C. arabica plants with these 4 QTNs showed qualitative resistance to BHB (Supplementary Table 3). Thirdly, the genes close to these QTNs are predicted to encode a serine/threonine-kinase protein and a CC-NBS-LRR protein, which act via an indirect interaction with the pathogen effector (Table 2). Finally, fourth, those genes showed a very similar transcriptional profile for cultivars Catuaí Vermelho IAC 99 (susceptible) and IAPAR 59 (quantitative resistance), while the cultivar IPR 102 (qualitative resistance) showed a different profile (Figures 5A, B), with these genes being upregulated in relation to the susceptible cultivars.

In Arabidopsis and tomato interaction with P. syringae the effector AvrPto binds to pattern recognition receptors (PRRs), including receptor-like protein kinases (RLKs) protein, and blocks pattern triggered immunity (PTI) activation. This block enhances host susceptibility. In resistant plants, a serine/threonine-kinase can mimic the kinase domain of the RLKs and bind to AvrPto. An NB-LRR protein recognizes the bind between the effector and the serine/threonine-kinase and then activates the effector triggered immunity (ETI). Thus, PTI and ETI are activated, resulting in a hypersensitivity response (HR) (Supplementary Figure 9) (Jones and Dangl, 2006; Zipfel and Rathjen, 2008; Zong et al., 2008).

Thus, the genes g000 (Chr_0_434_435) and g010741 (Chr_2_sg_E) could probably play a similar role in resistance to BHB in C. arabica. These results are an indication that the transcription levels for genes g000 (Chr_0_434_435) serine/threonine-kinase and g010741(Chr_2_sg_E) CC-NBS-LRR, found in IPR 102, may be necessary to activate qualitative resistance to BHB.

The haplotype analysis performed in this work identified a high number of polymorphisms in the CC-NBS-LRR gene, among C. arabica genotypes. Interestingly, Et039 (Hap1), showed the alternative alleles associated with resistance to BHB for the QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G. Meanwhile, Caturra and Geisha (Hap2) showed the standard alleles for the same QTNs (Figure 6C). The nsSNP identified between Hap1 and Hap2, mainly those inserted in conserved domains (Figure 6B), may be related to the defense response functionality of the gene. However, the haplotypes must be validated in a greater number of genotypes with phenotyping for resistance to BHB in order to reinforce the role of this genes in the resistance to BHB.

Relationship between the loci involved in resistance to BHB and the C. arabica domestication process

The population structure results showed by hierarchical analysis, PCA, and Bayesian approach at K = 3 were consistent for clustering the C. arabica Ethiopian accessions and the cultivars according to their domestication level characteristics. The admixed accessions identification was possible with the Bayesian approach. Furthermore, the Bayesian approach at K = 3 was the most suitable in demonstrating the genetic proximity between the accessions collected at the East of the Great Rift Valley and the C. arabica cultivars. This result corroborate with other studies that identified the movement of C. arabica from West to East of the Great Rift Valley, in direction to Yemen, from where C. arabica was disperse worldwide (Montagnon and Bouharmont, 1996; Silvestrini et al., 2007; Scalabrin et al., 2020).

From the frequency analysis of resistance to BHB into the clusters formed in population structure Bayesian approach at K = 3, it was possible to identify a higher frequency of resistance to BHB in the “domesticated” cluster (Figure 7A). However, in the C. arabica cultivars, there was a decrease in the frequency of accessions resistant to BHB. The QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G, with the highest -log10 (p-values) of association identified by both the single-locus and multi-locus methods, showed a higher frequency in the “domesticated” cluster (Figure 7B). The distribution frequency of the others QTNs associated with resistance to BHB identified in this work are shown in Supplementary Figure 7.

FIGURE 7
www.frontiersin.org

Figure 7 Population structure and relation with resistance to BHB. (A) The distribution frequency of accessions resistance to BHB into the “cultivars” (n = 16), “domesticated” (n = 20), and “wild” (n = 33) clusters identified in population structure Bayesian approach at K = 3. C. arabica Ethiopian accessions classified as admixed were disregarded in this analysis. The resistance to BHB is based in a scale with a score from 0 to 5, where 0 corresponds to highly resistant plants, 1 to resistant, 2 to moderately resistant, 3 to moderately susceptible, 4 to susceptible, and 5 to highly susceptible. (B) The distribution frequency of the QTNs associated with resistance to BHB into the “cultivars”, “domesticated”, and “wild” clusters. QTN associated with resistance to BHB with its respective location in the C. arabica Et039 genome1; QTN effect2; r2 (%) of the QTN in relation to the resistance phenotype3.

The highest frequency, in the “domesticated” cluster, of C. arabica accessions resistant to BHB and of the QTNs Chr_2_sg_E_32049720_G and Chr_2_sg_E_32049728_G, suggest that resistance to BHB is a trait inserted in loci that may have had an influence in plants selection during the C. arabica domestication history (Figures 7A, B). The influence of genomic regions linked to resistance to disease in the domestication process has already been described for other species, such as peanut (Arachis hypogaea), tomato (Solanum pimpinellifolium L. and S. lycopersicum L.), and soybean (Glycine max L.) (Zhao et al., 2015; Zhang et al., 2017; Razifard et al., 2020). In addition, genomic regions of resistance to disease showed the highest selection rate, during the divergence between Coffea arabica, C. canephora, and C. excelsa (Huang et al., 2020).

However, the lower frequency in the “cultivar” cluster of both C. arabica resistant to BHB and QTNs with the highest -log10 (p-values) identified in this GWAS agrees with the low genetic variability and low resistance to pathogens known in C. arabica cultivars (van der Vossen, 1985; Anthony et al., 2002). The lack of resistance to diseases in cultivars, a disadvantage in relation to the diversity found in wild genotypes, is extensively reported in the literature in many different species (Lam et al., 2011; Stalker et al., 2013; Cao et al., 2014). According to these studies, characteristics such as greater palatability, greater production, and fruit size, had greater relevance in the cultivar development over resistance to diseases. In the case of C. arabica, most cultivars are derived from only two varieties: Typica and Bourbon (Krug et al., 1939). This fact makes the genetic diversity among commercial C. arabica cultivars quite limited (Anthony et al., 2002; Steiger et al., 2002; Vidal et al., 2010). In addition, breeding programs that originated these cultivars prioritized criteria such as high productivity, beverage quality, and drought tolerance (van der Vossen, 1985). As a result, a large genetic bottleneck was formed. It is important to observe that, in this work, the haplotype analysis of a CC-NBS-LRR gene showed a greater genetic diversity from an Ethiopian accession (C. arabica Et039) than in commercial genotypes (Caturra and Geisha).

These results reinforce the importance of conservation of C. arabica wild types, both in their native habitats and in germplasm banks. Since their genetic diversity can be used to identify genes involved in traits of interest, such as resistance to diseases, and to elucidate the C. arabica domestication process.

Conclusions

This study provides 11 QTNs associated with resistance to BHB that have potential for application in Marker Assisted Selection and the characterization of a serine/threonine-kinase and a CC-NBS-LRR, probably, involved in qualitative resistance to BHB, could help to direct the breeding and could be a gene editing target.

Our results demonstrate that the combination of different GWAS methods can increase the efficiency and accuracy of obtaining a panel of QTNs associated with features of interest.

Furthermore, this work provides an insight into the relationship between the loci involved in resistance to BHB and the C. arabica domestication history. This reinforces the importance of maintenance of the Ethiopian coffee collections as a genetic diversity source for breeding programs on the pathway to achieve sustainable agriculture.

This is the first report for the interaction between C. arabica and P. syringae at the genetics and molecular level. In this way, the results presented are a starting point for future works that may bring greater evidence and validation of the resistance genes identified, such as gene expression analysis and validation in gene knock out mutants or gene overexpressing mutants.

Data availability statement

The GBS data presented in the study are deposited in the National Center for Biotechnology Information repository, accession number PRJNA882422.

Author contributions

CA, LP wrote the manuscript. CA, GCS, LFPP designed the experiment. MSF, RVF, LMN, BSS performed genotyping data. CA, GHS, LMRR, SALD participated in preparation of the materials and phenotype analysis. DSD, LFPP, MLVR provide funds for research. All the authors read, contributed and approved the final version of the manuscript. All authors contributed to the article and approved the submitted version.

Funding

This work was supported by grants from the Consórcio Pesquisa Café (Grant 10.18.20.027.00), INCT Café and FAPEMIG. CA, RVF, LMN, BSS and MSF received a PhD fellowship from CAPES (Finance Code 001). LP and DD received a Research Fellowship from CNPq (311452/2020-5).

Acknowledgments

We gratefully acknowledge the Arabica Coffee Genome Consortium (ACGC) for providing access to C. arabica Et039 genome data. We also acknowledge the Consórcio Pesquisa Café, INCT Café, Fundação de Amparo à Pesquisa do Estado de Minas Gerais (FAPEMIG), the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), and the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) for financial support and fellowships.

Conflict of interest

GCS was employed by company Tropical Melhoramento & Genética (TMG).

The remaining 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.2022.989847/full#supplementary-material

Supplementary Figure 1 | Histogram of the disease distribution, values of response to Bacterial Halo Blight obtained in field evaluation (Mohan et al., 1978; Ito et al., 2008). The X-axis represents the classes of distribution for the 120 C. arabica wild accessions (blue), 11 C. arabica cultivars (red) and BA-10 genotype evaluated. The Y-axis shows the count of C. arabica genotypes in each category.

Supplementary Figure 2 | Distribution of 11,290 SNPs, identified in 120C. arabica wild accessions, 11 C. arabica cultivars and BA-10 genotype, in the C. arabica genome. Each chromosome has its size information in base pairs. The bars correspond to the number of SNPs on each chromosome.

Supplementary Figure 3 | Linkage disequilibrium (LD) decay for 16,146 SNPs obtained after quality filters and the removal of SNPs from Chr zero. Each point represents a physical distance (bp) between SNPs pairs and their squared correlation coefficient (r2). The red line represents the line for non-linear regression against physical distance. The green line represents the LD decay value for r2 = 0.2.

Supplementary Figure 4 | Dissimilarity matrix representing the genetic distance between the 120 C. arabica wild accessions, 11 C. arabica cultivars and BA-10 genotype. The color intensity is proportional to the value of genetic distance between the accessions. The matrix was constructed based on the 11,290 SNPs identified in this population. The color intensity is proportional to the value of genetic distance between the accessions.

Supplementary Figure 5 | Structure analysis of 120 C. arabica wild accessions, 11 C. arabica cultivars and BA-10 genotype. Evolution of ΔK values (y-axis) according to the number of genetic groups (x-axis).

Supplementary Figure 6 | Classification of 132 accessions of C. arabica, using 11,290 SNPs identified in this population, into K = 2, K = 3, K = 4, and K = 5 using STRUCTURE 2.3.4. The distribution of the accessions to different populations is indicated by the color code. Numbers on the y-axis show the subgroup membership, and the x-axis shows the different accessions.

Supplementary Figure 7 | Population structure and relation with resistance to BHB. The distribution frequency of the QTNs associated with resistance to BHB into the “cultivars” (n = 16), “domesticated” (n = 20), and “wild” (n = 33) clusters identified in population structure Bayesian approach at K = 3. C. arabica Ethiopian accessions classified as admixed were disregarded in this analysis. QTN associated with resistance to BHB with its respective location in the C. arabica Et039 genome1; QTN effect2; r2 (%) of the QTN in relation to the resistance phenotype3.

Supplementary Figure 8 | Transcribed levels in the absence of the pathogen. (A) The g000 (Chr_0_434_435) and (B) g010741 (Chr_2_sg_E). Data represent the means ±  standard deviation. Upper case letters compare columns of different genotypes. Same letters show no difference by Tukey test (for P ≤ 0.05).

Supplementary Figure 9 | Model for PTI-inhibition and ETI-induction by Pseudomonas syringae effector. In susceptible plants, the effectors bind to PRRs. This blocks PTI activation and enhances host susceptibility to the bacterium. In resistant plants, serine/threonine – kinase is evolved to mimic the kinase domain of PRRs and binds to the effectors to trigger ETI, resulting in a resistance outcome. Blocking (A) and triggering (B) a hypersensitive response by Pseudomonas syringae effector.

References

Adugna, G., Jefuka, C., Teferi, D., Abate, S. (2012). New record and outbreaks of bacterial blight of coffee (Pseudomonas syringae) in southern Ethiopia: impact of climate change scenarios. in 24th international conference on coffee science (ASIC) São José, Costa Rica, 85.

Google Scholar

Amaral, J. F., Teixeira, C., Pinheiro, E. D. (1956). A bactéria causadora da mancha-aureolada do cafeeiro. Arquivo do Instituto. Biológico SP Brazil 23, 151–155.

Google Scholar

Andersen, E. J., Ali, S., Byamukama, E., Yen, Y., Nepal, M. P. (2018). Disease resistance mechanisms in plants. Genes 9, 339–368. doi: 10.3390/genes9070339

CrossRef Full Text | Google Scholar

Anthony, F., Bertrand, B., Quiros, O., Wilches, A., Lashermes, P., Berthaud, J., et al. (2001). Genetic diversity of wild coffee (Coffea arabica l.) using molecular markers. Euphytica 118, 53–65. doi: 10.1023/A:1004013815166

CrossRef Full Text | Google Scholar

Anthony, F., Combes, M. C., Astorga, C., Bertrand, B., Graziosi, G., Lashermes, P. (2002). The origin of cultivated Coffea arabica l. varieties revealed by AFLP and SSR markers. Theor. Appl. Genet. 104, 894–900. doi: 10.1007/s00122-001-0798-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Anzueto, F., Bertrand, B., Sarah, J. L., Eskes, A. B., Decazy, B., Rica, C., et al. (2001). Resistance to Meloidogyne incognita in Ethiopian Coffea arabica accessions. Euphytica 118, 1–8. doi: 10.1023/A:1003712232325

CrossRef Full Text | Google Scholar

Bachman, E. S., McClay, D. R. (1996). Molecular cloning of the first metazoan beta-1,3 glucanase from eggs of the sea urchin Strongylocentrotus purpuratus. Proc. Natl. Acad. Sc.i U S A. 93 (13), 6808–6813. doi: 10.1073/pnas.93.13.6808

CrossRef Full Text | Google Scholar

Balasubramanian, V., Vashisht, D., Cletus, J., Sakthivel, N. (2012). Plant β-1,3-glucanases: their biological functions and transgenic expression against phytopathogenic fungi. Biotechnol. Lett. 34 (11), 1983–1990. doi: 10.1007/s10529-012-1012-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Balmant, K., Parker, J., Yoo, M. J., Zhu, N., Dufresne, C., Chen, S. (2015). Redox proteomics of tomato in response to Pseudomonas syringae infection. Hortic. Res. 2, 15043. doi: 10.1038/hortres.2015.43

PubMed Abstract | CrossRef Full Text | Google Scholar

Barsalobres-Cavallari, C. F., Severino, F. E., Maluf, M. P., Maia, I. G. (2009). Identification of suitable internal control genes for expression studies in Coffea arabica under different experimental conditions. BMC Mol. Biol. 10, 1. doi: 10.1186/1471-2199-10-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Bertrand, B., Vaast, P., Alpizar, E., Etienne, H., Davrieux, F., Charmetant, P. (2006). Comparison of bean biochemical composition and beverage quality of arabica hybrids involving Sudanese-Ethiopian origins with traditional varieties at various elevations in central America. Tree Physiol. 26, 1239–1248. doi: 10.1093/treephys/26.9.1239

PubMed Abstract | CrossRef Full Text | Google Scholar

Bittner-Eddy, P. D., Crute, I. R., Holub, E. B., Beynon, J. L. (2000). RPP13 is a simple locus in Arabidopsis thaliana for alleles that specify downy mildew resistance to different avirulence determinants in Peronospora parasitica. Plant J. 21, 177–188. doi: 10.1046/j.1365-313x.2000.00664.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Blum, M., Chang, H. Y., Chuguransky, S., Grego, T., Kandasaamy, S., Mitchell, A., et al. (2021). The InterPro protein families and domains database: 20 years on. Nucleic Acids Res. 49 (D1), 344–354. doi: 10.1093/nar/gkaa977

CrossRef Full Text | Google Scholar

Boller, T., Felix, G. (2009). A renaissance of elicitors: perception of microbe-associated molecular patterns and danger signals by pattern-recognition receptors. Annu. Rev. Plant Biol. 60, 379–406. doi: 10.1146/annurev.arplant.57.032905.105346

PubMed Abstract | CrossRef Full Text | Google Scholar

Bove, J., Kim, C. Y., Gibson, C. A., Assmann, S. M. (2008). Characterization of wound-responsive RNA-binding proteins and their splice variants in arabidopsis. Plant Mol. Biol. 67 (1-2), 71–88. doi: 10.1007/s11103-008-9302-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Boyd, L. A., Ridout, C., Sullivan, D. M. O., Leach, J. E., Leung, H. (2012). Plant – pathogen interactions: Disease resistance in modern agriculture. Trends Genet. 29, 233–240. doi: 10.1016/j.tig.2012.10.011

PubMed Abstract | CrossRef Full Text | Google Scholar

Bradbury, P. J., Zhang, Z., Kroon, D. E., Casstevens, T. M., Ramdoss, Y., Buckler, E. S. (2007). TASSEL: Software for association mapping of complex traits in diverse samples. Bioinformatics 23, 2633–2635. doi: 10.1093/bioinformatics/btm308

PubMed Abstract | CrossRef Full Text | Google Scholar

Brun, H., Chèvre, A., Fitt, B. D. L., Powers, S., Besnard, A., Ermel, M., et al. (2010). Quantitative resistance increases the durability of qualitative resistance to Leptosphaeria maculans in Brassica napus. New Phytol. 185, 285–299. doi: 10.1111/j.1469-8137.2009.03049.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Cao, K., Zheng, Z., Wang, L., Liu, X., Zhu, G., Fang, W., et al. (2014). Comparative population genomics reveals the domestication history of the peach, Prunus persica, and human influences on perennial fruit crops. Genome Biol. 15, 415–429. doi: 10.1186/s13059-014-0415-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Chang, S., Puryear, J., Cairney, J. (1993). A simple and efficient method for isolating RNA from pine trees. Plant Mol. Biol. 11, 113–116. doi: 10.1007/BF02670468

CrossRef Full Text | Google Scholar

Cheval, C., Aldon, D., Galaud, J., Ranty, B. (2013). Calcium/calmodulin-mediated regulation of plant immunity. Biochim. Biophys. Acta 1833, 1766–1771. doi: 10.1016/j.bbamcr.2013.01.031

PubMed Abstract | CrossRef Full Text | Google Scholar

Clarindo, W. R., Carvalho, C. R. (2008). First Coffea arabica karyogram showing that this species is a true allotetraploid. Plant Syst. Evol. 274, 237–241. doi: 10.1007/s00606-008-0050-y

CrossRef Full Text | Google Scholar

Coll, N. S., Epple, P., Dangl, J. L. (2011). Programmed cell death in the plant immune system. Cell Death Differ. 18, 1247–1256. doi: 10.1038/cdd.2011.37

PubMed Abstract | CrossRef Full Text | Google Scholar

Corwin, J. A., Kliebenstein, D. J. (2017). Quantitative resistance: more than just perception of a pathogen. Plant Cell 29, 655–665. doi: 10.1105/tpc.16.00915

PubMed Abstract | CrossRef Full Text | Google Scholar

Cui, H., Tsuda, K., Parker, J. E. (2015). Effector-triggered immunity: from pathogen perception to robust defense. Annu. Rev. Plant Biol. 66, 487–511. doi: 10.1146/annurev-arplant-050213-040012

PubMed Abstract | CrossRef Full Text | Google Scholar

Cui, Y., Zhang, F., Zhou, Y. (2018). The application of multi-locus GWAS for the detection of salt-tolerance loci in rice. Front. Plant Sci. 9. doi: 10.3389/fpls.2018.01464

CrossRef Full Text | Google Scholar

Davis, A. P., Gole, T. W., Baena, S., Moat, J. (2012). The impact of climate change on indigenous arabica coffee (Coffea arabica): predicting future trends and identifying priorities. Plos One 7, e47981. doi: 10.1371/journal.pone.0047981

PubMed Abstract | CrossRef Full Text | Google Scholar

Davis, A. P., Tosh, J., Ruch, N., Fay, M. (2011). Growing coffee: Psilanthus (Rubiaceae) subsumed on the basis of plastid and nuclear DNA sequences; implication for the size, morphology, distribution and evolutionary history of coffea. Bot. J. Linn. Soc 167, 357–377. doi: 10.1111/j.1095-8339.2011.01177.x

CrossRef Full Text | Google Scholar

Ding, R., Yang, M., Quan, J., Li, S., Zhuang, Z., Zhou, S., et al. (2019). Single-locus and multi-locus genome-wide association studies for intramuscular fat in duroc pigs. Front. Plant Sci. 10. doi: 10.3389/fgene.2019.00619

CrossRef Full Text | Google Scholar

Earl, D. A., von Holdt, B. M. (2012). Structure harvester: A website and program for visualizing STRUCTURE output and implementing the evanno method. Conserv. Genet. Resour. 4, 359–361. doi: 10.1007/s12686-011-9548-7

CrossRef Full Text | Google Scholar

Edgar, R. C. (2004). MUSCLE: a multiple sequence alignment method with reduced time and space complexity. BMC Bioinform. 5, 113. doi: 10.1186/1471-2105-5-113

CrossRef Full Text | Google Scholar

Evanno, G., Regnaut, S., Goudet, J. (2005). Detecting the number of clusters of individuals using the software STRUCTURE: A simulation study. Mol. Ecol. 14, 2611–2620. doi: 10.1111/j.1365-294X.2005.02553.x

PubMed Abstract | CrossRef Full Text | Google Scholar

FAO: Food and Agriculture Organization (1968). Coffee mission to Ethiopia 1964–65 (Rome: FAO).

Google Scholar

Fatobene, B. J., dos, R., Andrade, V. T., Aloise, G. S., Silvarolla, M. B., Gonçalves, W., et al. (2017). Wild Coffea arabica resistant to Meloidogyne paranaensis and genetic parameters for resistance. Euphytica 213, 196. doi: 10.1007/s10681-017-1986-1

CrossRef Full Text | Google Scholar

Felicio, M. S. (2020). Estudo de associação genômica ampla aplicada ao conteúdo de macronutrientes em grãos de coffea arabica l. [doctor’s thesis] ([Botucatu (SP)]: Universidade Estadual Paulista).

Google Scholar

Harrell, E. F., Jr., Dupont, C. (2020). “). hmisc: harrell miscellaneous,” in R package version 4.4-0 (CRAN.R-project.org/package=Hmisc).

Google Scholar

Holderbaum, M. M., Ito, D. S., Santiago, D. C., Harumi, L., Fernandes, L. E., Sera, G. H. (2020). Arabica coffee accessions originated from Ethiopia with resistance to nematode Meloidogyne paranaensis. Aust. J. Crop Sci. 14, 1209–1213. doi: 10.21475/ajcs.20.14.08.p1763

CrossRef Full Text | Google Scholar

Horikoshi, M., Tang, Y. (2016). “Ggfortify: data visualization tools for statistical analysis results,” in R package version 4.4-0 (CRAN.R-project.org/package=ggfortify).

Google Scholar

Huang, L., Wang, X., Dong, Y., Long, Y., Hao, C., Yan, L., et al. (2020). Resequencing 93 accessions of coffee unveils independent and parallel selection during Coffea species divergence. Plant Mol. Biol. 103, 51–61. doi: 10.1007/s11103-020-00974-4

PubMed Abstract | CrossRef Full Text | Google Scholar

Internacional Coffee Organization (2020) Trade statistics tables 2020. Available at: https://www.ico.org/prices/po-production.pdf (Accessed June 15, 2021).

Google Scholar

Ito, D. S., Sera, T., Sera, G. H., Del Grossi, L., Kanayama, F. S. (2008). Resistance to bacterial blight in arabica coffee cultivars. Crop Breed. Appl. Biotechnol. 8, 99–103. doi: 10.12702/1984-7033.v08n02a01

CrossRef Full Text | Google Scholar

Jones, J. D. G., Dangl, J. L. (2006). The plant immune system. Nat. Rev. 444, 323–329. doi: 10.1038/nature05286

CrossRef Full Text | Google Scholar

Jones, J. D. G., Vance, R. E., Dangl, J. L. (2016). Intracellular innate immune surveillance devices in plants and animals. Science 354, 1117–1126. doi: 10.1126/science.aaf6395

CrossRef Full Text | Google Scholar

Kassambara, A., Mundt, F. (2020). “Factoextra: extract and visualize the results of multivariate data analyses,” in R package version 1.0.7 (CRAN.R-project.org/package=factoextra).

Google Scholar

Kim, C. Y., Bove, J., Assmann, S. M. (2008). Overexpression of wound-responsive RNA-binding proteins induces leaf senescence and hypersensitive-like cell death. New Phytol. 180 (1), 57–70. doi: 10.1111/j.1469-8137.2008.02557.x

PubMed Abstract | CrossRef Full Text | Google Scholar

Krug, C. A., Mendes, J. E. T., Carvalho, A. (1939). “Taxonomia de coffea arabica l,” in Boletim técnico n°C62 (SP, Brazil: Instituto Agronômico de Campinas), 9–57.

Google Scholar

Kushalappa, A. C., Gunnaiah, R. (2013). Metabolo-proteomics to discover plant biotic stress resistance genes. Trends Plant Sci. 18, 522–531. doi: 10.1016/j.tplants.2013.05.002

PubMed Abstract | CrossRef Full Text | Google Scholar

Kushalappa, A. C., Yogendra, K. N., Karre, S. (2016). Plant innate immune response: qualitative and quantitative resistance. Crit. Rev. Plant Sci. 35, 38–55. doi: 10.1080/07352689.2016.1148980

CrossRef Full Text | Google Scholar

Lam, H., Xu, X., Liu, X., Chen, W., Yang, G., Wong, F., et al. (2011). Resequencing of 31 wild and cultivated soybean genomes identifies patterns of genetic diversity and selection. Nat. Genet. 42, 1053–1059. doi: 10.1038/ng.715

CrossRef Full Text | Google Scholar

Lashermes, P., Andrzejewski, S., Bertrand, B., Combes, M.-C., Dussert, S., Graziosi, G., et al. (2000). Molecular analysis of introgressive breeding in coffee (Coffea arabica l.). Theor. Appl. Genet. 100, 139–146. doi: 10.1007/s001220050019

CrossRef Full Text | Google Scholar

Lashermes, P., Combes, M. C., Robert, J., Trouslot, P., D’Hont, A., Anthony, F., et al. (1999). Molecular characterization and origin of the Coffea arabica l. genome. Mol. Gen. Genet. 261, 259–266. doi: 10.1007/s004380050965

PubMed Abstract | CrossRef Full Text | Google Scholar

Lewis, J. D., Lee, A. H. Y., Hassan, J. A., Wan, J., Hurley, B., Jhingree, J. R., et al. (2013). The Arabidopsis ZED1 pseudokinase is required for ZAR1-mediated immunity induced by the Pseudomonas syringae type III effector HopZ1a. PNAS 110 (46), 18722–18727. doi: 10.1073/pnas.131552011

PubMed Abstract | CrossRef Full Text | Google Scholar

Li, C., Fu, Y., Sun, R., Wang, Y., Wang, Q. (2018). Single-locus and multi-locus genome-wide association studies in the genetic dissection of fiber quality traits in upland cotton (Gossypium hirsutum l.). Front. Plant Sci. 9. doi: 10.3389/fpls.2018.01083

CrossRef Full Text | Google Scholar

Liscombe, D. K., Macleod, B. P., Loukanina, N., Nandi, O. I., Facchini, P. J. (2005). Evidence for the monophyletic evolution of benzylisoquinoline alkaloid biosynthesis in angiosperms. Phytochem. Lett. 66 (11), 1374–1393. doi: 10.1016/j.phytochem.2005.04.029

CrossRef Full Text | Google Scholar

Macho, A. P., Zipfel, C. (2015). Targeting of plant pattern recognition receptor-triggered immunity by bacterial type-III secretion system effectors. Curr. Opin. Microbiol. 23, 14–22. doi: 10.1016/j.mib.2014.10.009

PubMed Abstract | CrossRef Full Text | Google Scholar

Maciel, K. W., Destefano, S. A. L., Beriam, L. O. S., de Almeida, I. M. G., Patricio, F. R. A., Rodrigues, L. M. R., et al. (2018). Bacterial halo blight of coffee crop: aggressiveness and genetic diversity of strains. Bragantia. 77, 96–106. doi: 10.1590/1678-4499.2016267

CrossRef Full Text | Google Scholar

Maechler, M., Rousseeuw, P., Struyf, A., Hubert, M., Hornik, K. (2019). “Cluster: cluster analysis basics and extensions,” in R package version 2.1.0 (CRAN.R-project.org/package=cluster).

Google Scholar

Migicovsky, Z., Myles, S. (2017). Exploiting wild relatives for genomics-assisted breeding of perennial crops. Front. Plant Sci. 8. doi: 10.3389/fpls.2017.00460

CrossRef Full Text | Google Scholar

Misra, G., Badoni, S., Anacleto, R., Graner, A., Alexandrov, N., Sreenivasulu, N. (2017). Whole genome sequencing-based association study to unravel genetic architecture of cooked grain width and length traits in rice. Sci. Rep. 7, 12478. doi: 10.1038/s41598-017-12778-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Mohan, S. K., Cardoso, R. M. L., Paiva, M. A. (1978). Resistência em germoplasma de Coffea ao crestamento bacteriano incitado por Pseudomonas garcae. Pesquisa Agropecuária Bras. 13, 53–64.

Google Scholar

Money, D., Gardner, K., Migicovsky, Z., Schwaninger, H., Zhong, G. Y., Myles, S. (2015). LinkImpute: fast and accurate genotype imputation for nonmodel organisms. Genes Genomes Genet. 5, 2383–2390. doi: 10.1534/g3.115.021667

CrossRef Full Text | Google Scholar

Montagnon, C., Bouharmont, P. (1996). Multivariate analysis of phenotypic diversity of coffea arabica. Genet. Resour. Crop Evol. 43, 221–227. doi: 10.1007/BF00123274

CrossRef Full Text | Google Scholar

Moreno, A. A., Mukhtar, M. S., Blanco, F., Boatwright, J. L., Moreno, I., Jordan, M. R., et al. (2012). IRE1/bZIP60-mediated unfolded protein response plays distinct roles in plant immunity and abiotic stress responses. Plos One 7 (2), e319. doi: 10.1371/journal.pone.0031944

CrossRef Full Text | Google Scholar

Niks, R., Li, X., Marcel, T. C. (2015). Quantitative resistance to biotrophic filamentous plant pathogens: concepts, misconceptions, and mechanisms. Annu. Rev. Phytopathol. 53, 445–470. doi: 10.1146/annurev-phyto-080614-115928

PubMed Abstract | CrossRef Full Text | Google Scholar

Nobori, T., Wang, Y., Wu, J., Stolze, S. C., Tsuda, Y., Finkemeier, I., et al. (2020). Multidimensional gene regulatory landscape of a bacterial pathogen in plants. Nat. Plants 6, 883–896. doi: 10.1038/s41477-020-0690-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Pilet-Nayel, M.-L., Moury, B., Caffier, V., Montarry, J. (2017). Quantitative resistance to plant pathogens in pyramiding strategies for durable crop protection. Front. Plant Sci. 8. doi: 10.3389/fpls.2017.01838

CrossRef Full Text | Google Scholar

Pritchard, J. K., Stephens, M., Rosenberg, N. A., Donnelly, P. (2000). Association mapping in structured populations. Am. J. Hum. Genet. 67, 170–181. doi: 10.1086/302959

PubMed Abstract | CrossRef Full Text | Google Scholar

Queiroz-Voltan, R. B., Nardin, C. F., Fazuoli, L. C., Toma-Braghini, M. (2014). Caracterização da anatomia foliar de cafeeiros arábica em diferentes períodos sazonais. Biotemas 27, 1–10. doi: 10.5007/2175-7925.2014v27n4p1

CrossRef Full Text | Google Scholar

Quenouille, J., Paulhiac, E., Moury, B., Palloix, A. (2014). Quantitative trait loci from the host genetic background modulate the durability of a resistance gene : a rational basis for sustainable resistance breeding in plants. Heredity 112, 579–587. doi: 10.1038/hdy.2013.138

PubMed Abstract | CrossRef Full Text | Google Scholar

Ramos, A. H., Shavdia, L. D. (1976). A die-back of coffee in Kenya. Plant Dis. 60, 831–835.

Google Scholar

Razifard, H., Ramos, A., Della Valle, A. L., Bodary, C., Goetz, E., Elizabeth, J., et al. (2020). Genomic evidence for complex domestication history of the cultivated tomato in latin america. Mol. Biol. Evol. 37, 1118–1132. doi: 10.1093/molbev/msz297

PubMed Abstract | CrossRef Full Text | Google Scholar

Rodrigues, L. M. R., Almeida, I. M. G., Patrício, F. R. A., Beriam, L. O. S., Maciel, K. W., Braghini, M. T., et al. (2017b). Aggressiveness of strains and inoculation methods for resistance assessment to bacterial halo blight on coffee seedlings. J. Phytopathol. 165, 105–114. doi: 10.1111/jph.12543

CrossRef Full Text | Google Scholar

Rodrigues, L. M. R., Sera, G. H., Guerreiro Filho, O., Beriam, L. O. S., de Almeida, I. M. G. (2017a). First report of mixed infection by Pseudomonas syringae pathovars garcae and tabaci on coffee plantations. Plant Prot. 76, 543–549. doi: 10.1590/1678-4499.2016.399

CrossRef Full Text | Google Scholar

RStudio Team (2020). RStudio: integrated development for r. RStudio (Boston, MA: PBC). Available at: http://www.rstudio.com/.

Google Scholar

Rutledge, R. G., Stewart, D. (2008). Critical evaluation of methods used to determine amplification efficiency refutes the exponential character of real-time PCR. BMC Mol. Biol. 9, 96. doi: 10.1186/1471-2199-9-96

PubMed Abstract | CrossRef Full Text | Google Scholar

Salojärvi, J., Arabica Coffee Genome Consortium. (2021). Chromosome-level assembly of allotetraploid Coffea arabica reveals the complex history of a recent allopolyploid. 28th International Conference on Coffee Science (ASIC), Montpellier, France S1-PO-19.

Google Scholar

Sant’ana, G. C., Pereira, L. F. P., Pot, D., Ivamoto, S. T., Domingues, D. S., Ferreira, R. V., et al. (2018). Genome-wide association study reveals candidate genes influencing lipids and diterpenes contents in Coffea arabica l. Sci. Rep. 8, 465. doi: 10.1038/s41598-017-18800-1

PubMed Abstract | CrossRef Full Text | Google Scholar

Scalabrin, S., Toniutti, L., Di Gaspero, G., Scaglione, D., Magris, G., Vidotto, M., et al. (2020). A single polyploidization event at the origin of the tetraploid genome of Coffea arabica is responsible for the extremely low genetic variation in wild and cultivated germplasm. Sci. Rep. 10, 4642. doi: 10.1038/s41598-020-61216-7

PubMed Abstract | CrossRef Full Text | Google Scholar

Scholz, S. B. M., Kitzberger, C. S. G., Pagiatto, N. F., Pereira, L. F. P., Davrieux, F., Pot, D., et al. (2016). Chemical composition in wild ethiopian arabica coffee accessions. Euphytica 209, 429–438. doi: 10.1007/s10681-016-1653-y

CrossRef Full Text | Google Scholar

Silvarolla, M. B., Mazzafera, P., Fazuoli, L. C. (2004). A naturally decaffeinated arabica coffee. Nature 429, 826. doi: 10.1038/429826a

PubMed Abstract | CrossRef Full Text | Google Scholar

Silvestrini, M., Junqueira, M. G., Favarin, A. C., Guerreiro-Filho, O., Maluf, M. P., Silvarolla, M. B., et al. (2007). Genetic diversity and structure of Ethiopian, Yemen and Brazilian Coffea arabica l. accessions using microsatellites markers. Genet. Resour. Crop Evol. 54, 1367–1379. doi: 10.1007/s10722-006-9122-4

CrossRef Full Text | Google Scholar

Stalker, H. T., Tallury, S. P., Bertioli, D., Bertioli, S. C. L. (2013). The value of diploid peanut relatives for breeding and genomics. Peanut Sci. 40, 70–88. doi: 10.3146/PS13-6.1

CrossRef Full Text | Google Scholar

St. Clair, D. A. (2010). Quantitative disease resistance and quantitative resistance loci in breeding. Annu. Rev. Phytopathol. 48, 247–268. doi: 10.1146/annurev-phyto-080508-081904

PubMed Abstract | CrossRef Full Text | Google Scholar

Steiger, D. L., Nagai, C., Moore, P. H., Morden, C. W., Osgood, R. V., Ming, R. (2002). AFLP analysis of genetic diversity within and among Coffea arabica cultivars. Theor. Appl. Genet. 105, 209–215. doi: 10.1007/s00122-002-0939-8

PubMed Abstract | CrossRef Full Text | Google Scholar

Su, J., Ma, Q., Li, M., Hao, F., Wang, C. (2018). Multi-locus genome-wide association studies of fiber-quality related traits in chinese early-maturity upland cotton. Front. Plant Sci. 9. doi: 10.3389/fpls.2018.01169

CrossRef Full Text | Google Scholar

Tamba, C. L., Ni, Y. L., Zhang, Y. M. (2017). Iterative sure independence screening EM-Bayesian LASSO algorithm for multi-locus genome-wide association studies. Plos Comput. Biol. 13, e1005357. doi: 10.1371/journal.pcbi.1005357

PubMed Abstract | CrossRef Full Text | Google Scholar

Tamba, C. L., Zhang, Y. (2018). A fast mrMLM algorithm for multi-locus genome-wide association studies. BioRxiv 341784. doi: 10.1101/341784v1

CrossRef Full Text | Google Scholar

van der Vossen, H. A. M. (1985). ““Coffee selection and breeding”,” in Coffee: botany biochemistry and production of beans and beverage. Eds. Clifford, M. N., Wilson, K. C. (Westport, London: Croom Herm), 48–96.

Google Scholar

van der Vossen, H., Bertrand, B., Charrier, A. (2015). Next generation variety development for sustainable production of arabica coffee (Coffea arabica l.): a review. Euphytica 204, 243–256. doi: 10.1007/s10681-015-1398-z

CrossRef Full Text | Google Scholar

van der Vossen, H. A. M., Walyaro, D. J. (2009). Additional evidence for oligogenic inheritance of durable host resistance to coffee berry disease (Colletotrichum kahawae) in arabica coffee (Coffea arabica l.). Euphytica 165, 105–111. doi: 10.1007/s10681-008-9769-3

CrossRef Full Text | Google Scholar

Vidal, R. O., Mondego, J. M. C., Pot, D., Ambrosio, A. B., Andrade, A. C., Pereira, L. F. P., et al. (2010). A high-throughput data mining of single nucleotide polymorphisms in Coffea species expressed sequence tags suggests differential homeologous gene expression in the allotetraploid Coffea arabica. Plant Physiol. 154, 1053–1066. doi: 10.1104/pp.110.162438

PubMed Abstract | CrossRef Full Text | Google Scholar

Villa-Rivera, M. G., Cano-Camacho, H., López-Romero, E., Zavala-Páramo, M. G. (2021). The role of arabinogalactan type II degradation in plant-microbe interaction. Front. Microbiol. 12. doi: 10.3389/fmicb.2021.730543

CrossRef Full Text | Google Scholar

Wang, S. B., Feng, J. Y., Ren, W. L., Huang, B., Zhou, L., Wen, Y. J., et al. (2016). Improving power and accuracy of genome-wide association studies via a multi-locus mixed linear model methodology. Sci. Rep. 6, 19444. doi: 10.1038/srep19444

PubMed Abstract | CrossRef Full Text | Google Scholar

Wei, T., Simko, V. (2021). “R package “corrplot”: visualization of a correlation matrix (version 0.88),” in CRAN (R-project.org/package=corrplot).

Google Scholar

Wen, Y. J., Zhang, H., Ni, Y. L., Huang, B., Zhang, J., Feng, J. Y., et al. (2017). Methodological implementation of mixed linear models in multi-locus genome-wide association studies. Brief. Bioinform. 19, 700–712. doi: 10.1093/bib/bbw145

CrossRef Full Text | Google Scholar

Worku, M., De.Meulenaer, B., Duchateau, L. (2018). Effect ofaltitude on biochemical composition and quality of green arabica coffee beans can be affected by shade and postharvest processing method. Food Res. Int. 105, 278–285. doi: 10.1016/j.foodres.2017.11.016

PubMed Abstract | CrossRef Full Text | Google Scholar

Xuehui, B., Lihong, Z., Yongliang, H., Guanghai, J., Jinhong, L., Zhang, H. (2013). Isolation and identification of the pathogen of coffee bacterial blight disease. Chin. J. Trop. Crop 34, 738–742. doi: 10.3969/j.issn.1000-2561.2013.04.028

CrossRef Full Text | Google Scholar

Xu, Y., Yang, T., Zhou, Y., Yin, S., Li, P., Liu, J., et al. (2018). Genome-wide association mapping of starch pasting properties in maize using single-locus and multi-locus models. Front. Plant Sci. 9. doi: 10.3389/fpls.2018.01311

CrossRef Full Text | Google Scholar

Yogendra, K. N., Dhokane, D., Kushalappa, A. C., Sarmiento, F., Rodriguez, E., Mosquera, T. (2017). StWRKY8 transcription factor regulates benzylisoquinoline alkaloid pathway in potato conferring resistance to late blight. Plant Sci. 256, 208–216. doi: 10.1016/j.plantsci.2016.12.014

PubMed Abstract | CrossRef Full Text | Google Scholar

Yuan, M., Jiang, Z., Bi, G., Nomura, K., Liu, M., Wang, Y., et al. (2021). Pattern-recognition receptors are required for NLR-mediated plant immunity. Nature 592, 105–109. doi: 10.1038/s41586-021-03316-6

PubMed Abstract | CrossRef Full Text | Google Scholar

Yu, J., Pressoir, G., Briggs, W. H., Bi, I. V., Yamasaki, M., Doebley, J. F., et al. (2006). A unified mixed-model method for association mapping that accounts for multiple levels of relatedness. Nat. Genet. 38, 203–208. doi: 10.1038/ng1702

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, Y. M., Jia, Z., Dunwell, J. M. (2019). Editorial: The applications of new multi-locus gwas methodologies in the genetic dissection of complex traits. Front. Plant Sci. 10. doi: 10.3389/fpls.2019.00100

CrossRef Full Text | Google Scholar

Zhang, Q., Wu, L., Yin, H., Xu, Z., Zhao, Y., Gao, M., et al. (2021). D6 protein kinase in root xylem benefiting resistance to Fusarium reveals infection and defense mechanisms in tung trees. Hortic. Res. 8, 240. doi: 10.1038/s41438-021-00656-2

PubMed Abstract | CrossRef Full Text | Google Scholar

Zhang, X., Zhang, J., He, X., Wang, Y., Ma, X., Yin, D. (2017). Genome-wide association study of major agronomic traits related to domestication in peanut. Front. Plant Sci. 8. doi: 10.3389/fpls.2017.01611

CrossRef Full Text | Google Scholar

Zhao, S., Zheng, F., He, W., Wu, H., Pan, S., Lam, H. (2015). Impacts of nucleotide fixation during soybean domestication and improvement. BMC Plant Biol. 15, 81. doi: 10.1186/s12870-015-0463-z

PubMed Abstract | CrossRef Full Text | Google Scholar

Zipfel, C., Rathjen, J. P. (2008). Plant immunity: AvrPto targets the frontline. Curr. Biol. 18, 218–220. doi: 10.1016/j.cub.2008.01.016

CrossRef Full Text | Google Scholar

Zong, N., Xiang, T., Zou, Y., Chai, J., Zhou, J. M. (2008). Blocking and triggering of plant immunity by Pseudomonas syringae effector AvrPto. Plant Signal. Behav. 3, 583–585. doi: 10.4161/psb.3.8.5741

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: coffee domestication, bacterial halo blight, plant disease, single-locus genome-wide association, multi-locus genome-wide association

Citation: Ariyoshi C, Sant’ana GC, Felicio MS, Sera GH, Nogueira LM, Rodrigues LMR, Ferreira RV, da Silva BSR, Resende MLVd, Destéfano SAL, Domingues DS and Pereira LFP (2022) Genome-wide association study for resistance to Pseudomonas syringae pv. garcae in Coffea arabica. Front. Plant Sci. 13:989847. doi: 10.3389/fpls.2022.989847

Received: 08 July 2022; Accepted: 22 September 2022;
Published: 18 October 2022.

Edited by:

Elisabetta Mazzucotelli, Council for Agricultural and Economics Research (CREA), Italy

Reviewed by:

Yuan-Ming Zhang, Huazhong Agricultural University, China
Maria Celeste Gonçalves-Vidigal, Universidade Estadual de Maringá, Brazil

Copyright © 2022 Ariyoshi, Sant’ana, Felicio, Sera, Nogueira, Rodrigues, Ferreira, da Silva, Resende, Destéfano, Domingues and Pereira. 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: Luiz Filipe Protasio Pereira, filipe.pereira@embrapa.br

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.