Next Article in Journal
Pathogenic Mitochondria DNA Mutations: Current Detection Tools and Interventions
Next Article in Special Issue
Ectopic Expression of AhGLK1b (GOLDEN2-like Transcription Factor) in Arabidopsis Confers Dual Resistance to Fungal and Bacterial Pathogens
Previous Article in Journal
The Bacillus Subtilis K-State Promotes Stationary-Phase Mutagenesis via Oxidative Damage
Previous Article in Special Issue
Tomato Natural Resistance Genes in Controlling the Root-Knot Nematode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Identification and Validation of Candidate Genes Conferring Resistance to Downy Mildew in Maize (Zea mays L.)

1
Department of Life Science, Dongguk University-Seoul, Seoul 04620, Korea
2
Department of Plant Resources, College of Industrial Science, Kongju National University, Yesan 32439, Korea
*
Author to whom correspondence should be addressed.
Genes 2020, 11(2), 191; https://doi.org/10.3390/genes11020191
Submission received: 2 January 2020 / Revised: 5 February 2020 / Accepted: 7 February 2020 / Published: 11 February 2020
(This article belongs to the Special Issue Genetics and Epigenetics of Biotic Stress Response in Plants)

Abstract

:
Downy mildew (DM) is a major disease of maize that causes significant yield loss in subtropical and tropical regions around the world. A variety of DM strains have been reported, and the resistance to them is polygenically controlled. In this study, we analyzed the quantitative trait loci (QTLs) involved in resistance to Peronosclerospora sorghi (sorghum DM), P. maydis (Java DM), and Sclerophthora macrospora (crazy top DM) using a recombinant inbred line (RIL) from a cross between B73 (susceptible) and Ki11 (resistant), and the candidate genes for P. sorghi, P. maydis, and S. macrospora resistance were discovered. The linkage map was constructed with 234 simple sequence repeat (SSR) and restriction fragment length polymorphism (RFLP) markers, which was identified seven QTLs (chromosomes 2, 3, 6, and 9) for three DM strains. The major QTL, located on chromosome 2, consists of 12.95% of phenotypic variation explained (PVE) and a logarithm of odds (LOD) score of 14.12. Sixty-two candidate genes for P. sorghi, P. maydis, and S. macrospora resistance were obtained between the flanked markers in the QTL regions. The relative expression level of candidate genes was evaluated by quantitative real-time polymerase chain reaction (qRT-PCR) using resistant (CML228, Ki3, and Ki11) and susceptible (B73 and CML270) genotypes. For the 62 candidate genes, 15 genes were upregulated in resistant genotypes. Among these, three (GRMZM2G028643, GRMZM2G128315, and GRMZM2G330907) and AC210003.2_FG004 were annotated as leucine-rich repeat (LRR) and peroxidase (POX) genes, respectively. These candidate genes in the QTL regions provide valuable information for further studies related to P. sorghi, P. maydis, and S. macrospora resistance.

1. Introduction

Maize is the world’s leading cereal in terms of production, with 1093 million metric tons produced on 186 million hectares globally. Maize is grown in both temperate and tropical areas of the world and is largely (around 80%) produced under rainy conditions in sub-Saharan Africa, South and Southeast Asia, and Latin America. It is particularly susceptible to abiotic and biotic stress. Eight major countries growing maize (China, India, Indonesia, Nepal, Pakistan, the Philippines, Thailand, and Vietnam) produce 98% of Asia’s and 28% of the world’s maize yield. Heavy economic and yield losses have been recorded due to infection by downy mildew (DM) agents in the Philippines, Taiwan, Indonesia, Thailand, India, Japan, Australia, Venezuela, North America, Europe, West Africa, and other parts of the world [1,2,3,4,5,6].
DM is caused by obligate pathogens that cannot be cultured in the laboratory, and its sporulation prefers high relative humidity, a night temperature of 21–23 °C, and light drizzle with cool weather. It is spread by oospores that survive in the soil [7] and can be spread through infected seeds or from plant-to-plant by airborne conidia. Because of the systemic nature of DM, susceptible lines usually die when infected in the seedling emergence stage, and when plants are infected during later growth stages, they cannot develop the maize ear despite having survived. At least six pathogens that cause DM infection of maize in Asia have been reported, including sorghum DM (P. sorghi (Weston & Uppal)), Philippine DM (P. philippinensis (Weston) Shaw), Java DM (P. maydis (Raciborski)), sugarcane DM (P. sacchari (Miyabe) Shirai and Hara), brown stripe DM (Sclerophthora rayssiae var. zeae), crazy top DM (S. macrospora), and Rajasthan DM (P. heteropogoni) [2,6,8,9,10,11,12,13,14,15]. DM is widespread in tropical regions, although its origin is conjectural, and because of the diversity of the DM pathogens and their systemic nature, the development of resistant varieties is needed. Moreover, a renewed emphasis on cost-effectiveness and environmental safety that has brought about the application of DM management by the development of resistant varieties. According to studies on the interaction of maize and the pathogens, resistance to DM is polygenically controlled [7,13,16,17,18,19,20,21,22,23].
Quantitative trait locus (QTL) mapping enables the detection, localization, and characterization of genetic factors contributing to polygenically inherited variation [24]. In QTL studies, the use of recombinant inbred lines (RILs) has more advantages than F2 or backcross populations [25,26,27]. Furthermore, RILs have been used to identify QTLs for the European corn borer [28], thermotolerance [29], and grain yield [30] in maize.
Several groups have performed QTL mapping using diverse mapping populations. George et al. [2] reported six QTLs on five chromosomes (1, 2, 6, 7, and 10) in RILs from the cross Ki3 (resistant) × CML139 (susceptible) of advanced inbred lines from the International Maize and Wheat Improvement Center (CIMMYT) in Indonesia, the Philippines, and Thailand. A QTL on chromosome 6 at bin 6.05 was found to majorly affect resistance to five DM pathogens (P. sorghi, P. philippinensis, P. maydis, S. rayssiae var. zeae, and P. heteropogoni). Agrama et al. [31] reported three QTLs on chromosomes 1 and 9 utilizing RILs derived from a cross between G62 (resistant) and G58 (susceptible) Egyptian inbreds. Two QTLs on chromosome 1 had a minor effect and one on chromosome 9 had a major effect. Sabry et al. [32] reported three QTLs on chromosomes 2, 3, and 9 utilizing F3 in Egypt, Thailand, and southern Texas. One QTL on chromosome 2 had a major effect and the two on chromosomes 3 and 9 had minor effects. Other studies have reported QTLs for P. sorghi (sorghum DM) resistance on chromosomes 2, 3, 4, 5, 6, and 9 [21,33,34].
Considering the complexity of quantitative traits, these QTLs can be used for introgression by marker-assisted selection with further validation [33]. It is difficult to estimate disease reaction accurately because of the factors influencing DM, such as plant maturity and the amount of pathogen inoculum. If there are differences in the pathogen populations or environment by genotype interactions in different locations, the analysis of simple and accurately scored molecular markers for the resistance genes of DM could greatly benefit future efforts to prevent loss to disease [32]. In addition, the evaluation of several DM strains using a mapping population could contribute to the accurate assessment of genetic contributions to resistance.
Genome-wide comparative transcriptome analysis has been performed by using an RNA-seq method in cabbage [35], cucumber [36], grapevine [37], and pearl millet [38]. DM resistance has been identified using the mutants at DM genes in lettuce [39]. Interaction between genes and pathways related to resistance against powdery mildew (PM) in melon has been profiled through comparative transcriptome analysis [40]. In maize, most studies on DM resistance have been performed using QTL analysis. In addition, there is a lack of information on the candidate genes and pathways for DM resistance in maize.
The main objective of this study is to validate candidate genes for P. sorghi, P. maydis, and S. macrospora resistance obtained from the QTL information of DM generated by using 192 F7 families derived from the cross B73 × Ki11. Our approach to the location of the QTLs for resistance to P. sorghi, P. maydis, and S. macrospora in maize is based on restriction fragment length polymorphism (RFLP) and simple sequence repeat (SSR) markers. We obtained candidate genes for P. sorghi, P. maydis, and S. macrospora resistance near the flanked markers of the QTL positions. The candidate genes were validated via quantitative real-time polymerase chain reaction (qRT-PCR) using resistant (CML228, Ki3, and Ki11) and susceptible (B73 and CML270) genotypes and identified by Pfam analysis.

2. Materials and Methods

2.1. Plant Materials

The mapping population consisting of RILs from the cross B73 (susceptible) and Ki11 (resistant) was evaluated in this study. Ki11 is the DM-resistant parent from the yellow flint line with late maturity. This inbred line is derived from Suwan1, which was developed for resistance to sorghum DM (P. sorghi) at Kasetsart University in Bangkok, Thailand [41,42]. B73 was the DM-susceptible parent from the temperate stiff-stalk maize line. This inbred line was developed at Iowa State University (Ames, IA) in the United States and is known to be susceptible to DM and drought, as well as being photoperiod-sensitive [43,44,45,46]. The F1 families were planted in 2013, and then F7 families of the RILs (n = 192) were obtained by selfing each individual plant in an experimental field at Dongguk University, Korea in 2017.

2.2. Evaluation of the RILs for DM

Field experiments with P. sorghi (sorghum DM), P. maydis (Java DM), and S. macrospora (crazy top DM) were conducted for genotypes (B73, CML228, CML270, Ki3, and Ki11) in Phnom Penh, Cambodia in April and September 2015. The sets of F7 RILs and parents (B73 and Ki11) were screened in in Phnom Penh, Cambodia in September to November 2017 (11°27′08.5″ N 105°09′57.6″ E; high humidity averages of 96.1%, 94.6%, and 93.4% and low humidity averages of 58.6%, 60.9%, and 63.1% in September, October, and November, respectively). DM disease was scored in the field using a modified spreader-row technique [2]. Susceptible genotypes (B73 and CML270) were used as two 5-m-long spreader rows with 20 plants per row after every 10th row of the test entries on the border rows of the experimental block (Figure S1) [44,47]. Seeds of the spreader genotypes germinated after 7 days, then a DM-infected maize plant in a pot was placed in each experimental block for 2 weeks to infect the spreader plants. After removing the DM-infected maize plant pots, RILs were planted with 50 × 25 cm spacing between the spreader rows. After germination of the RILs, DM incidence was assessed every 7 days for 6 weeks by scoring for systemic infection [48,49]:
DM incidence (%) = (Number of infected plants/Total number of plants) × 100,
The results are categorized as 0% (no symptoms) = highly resistant (HR), 1–10% = resistant (R), 11–25% = moderately resistant (MR), 26–50% = moderately susceptible (MS), 51–75% = susceptible (S), and 76–100% = highly susceptible (HS) [44].
Spreader genotypes were confirmed as having been 100% infected with DM, and the percentage of disease incidence was determined. The resistant plants did not show systemic symptoms of DM (emergence of characteristic chlorotic leaves) [50].

2.3. DNA Preparation

Genomic DNA was extracted from the parent plants and 192 F7 individuals at the four-leaf stage using a modified DNA isolation protocol. The quality and quantity of the genomic DNA were analyzed with 1% agarose gel electrophoresis and Spectrophotometer (Model MN-913, Maestrogen, Hsinchu, Taiwan).

2.4. Molecular Marker Assay

A total of 727 SSR and RFLP markers covering all 10 chromosomes were obtained from the Maize Genetics and Genomics Database (MaizeGDB) and publications. The amplification was performed using a Takara PCR Thermal Cycler Dice Touch (Takara, Shiga, Japan) with 25 ng of gDNA and Takara Taq polymerase (Takara, Shiga, Japan). The polymerase chain reaction (PCR) conditions consisted of one cycle of 5 min at 94 °C for the initial denaturation and 30 cycles of 1 min at 94 °C, 1 min at 51–65 °C (based on the annealing temperatures standardized for different primers), and 30 s at 72 °C, with a final extension step for 10 min at 72 °C. The PCR products were separated by electrophoresis on 3% agarose gel with 1× Tris-acetate-ethylenediaminetetraacetic acid buffer at 100 V for 1 h and visualized using ethidium bromide staining. Markers were screened for polymorphism between the parent lines of the RILs. Of the 727 markers, 234 were found to be polymorphic based on agarose gel electrophoresis; these were used for genotyping the 192 RIL individuals (Table S1). Alleles of the parents of B73 and Ki11 were designated as 2 and 0, respectively.

2.5. Linkage Map Construction and QTL Analysis for P. sorghi, P. maydis, and S. macrospora Resistance

QTL IciMapping v. 4.1 software from the Quantitative Genetics Group was used for both linkage map construction and QTL analysis using the MAP and BIP functions, respectively [51,52]. For linkage map construction, groups were ordered with an LOD score of 3.0 and the nearest neighbor combined with the two-opt algorithm (nnTwoOpt). Rippling was ordered with the sum of adjacent recombination frequencies with a window size of 5. Recombination frequencies were transformed into cM distances between linked loci using Kosambi’s [53] mapping function. For QTL analysis, inclusive composite interval mapping of the additive (ICIM-ADD) function was applied with 1.0 step and 0.001 probability in the stepwise regression. The data of missing phenotypes were ordered with the Deletion command. The threshold LOD scores were calculated using 1,000 permutations with a type I error of 0.05 [51,54]. Suggestive QTLs with an average LOD value >3.0 in a dataset were noted. A QTL with an average LOD value > 3.0 and average phenotypic variance contribution > 10% was defined as a major QTL [55,56]. The QTLs in the present study were compared with previously published ones [2,21,31,32]. This software program uses an improved algorithm of composite interval mapping with increased power to detect QTLs to reduce false detection rates and create less biased QTL effect estimates by employing stepwise regression followed by QTL scanning [51].

2.6. RNA Extraction and Candidate Gene Screening with Quantitative Real-Time PCR (qRT-PCR)

Total RNA of healthy (control) and DM-infected parent lines was extracted from 6-week-old leaves of B73, CML228, CML270, Ki3, and Ki11 using TRIzol (Invitrogen, Carlsbad, CA, USA). For the qRT-PCR, cDNA was synthesized using ReverTra Ace® qPCR RT Master Mix with gDNA Remover (Toyobo, Osaka, Japan) using 1 μg of total RNA extracted from leaves.
We searched for annotated genes near 14 flanked markers in the reference genome of B73 RefGen_v2 (MaizeGDB 2019). Partially, annotated genes were reported to be associated with biotic and abiotic stress. We obtained 62 genes in seven QTL locations; these were spread over chromosomes 2, 3, 6, and 9. The information and predicted protein of candidate genes were obtained from maizeGDB (B73 RefGen_v2). In addition, 19 DM-related genes from Arabidopsis thaliana, Oryza sativa, and Zea mays were obtained from the NCBI database (2019). The qRT-PCR primer sets were designed based on the 62 candidate genes and 19 DM-related genes using Primer3 [57] to validate the expression value of each candidate genes compared to control and DM-infected maize (Tables S2 and S3). The qRT-PCR was performed by using a gene-specific primer set, the CFX96 Touch Real-Time PCR Detection System (Bio-Rad Laboratories, USA). The qRT-PCR mixture contained 20 ng of cDNA, 0.2 µM of each gene-specific primer, 8 µL of sterile water, and 10 µL of TOPreal ™ qPCR 2× PreMIX (containing SYBR Green with low ROX; Enzynomics, South Korea) in a total volume of 20 μL. The qRT-PCR conditions were 10 min at 95 °C for the initial denaturation, then 45 cycles of 10 s at 95 °C, 10 s at 55–60 °C (based on the annealing temperatures standardized for different primers), and 30 s at 72 °C. The maize 18s rRNA (AF168884.1), UBCP (GRMZM2G102471_T01), MEP (GRMZM2G018103_T01), and LUG (GRMZM2G425377_T01) were used as reference genes (Table S4) [58]. The specificity and efficiency of the amplicon were confirmed through the melting curve analysis from 65 to 95 °C after each qRT-PCR. Three replicates of each experiment were performed for each candidate gene. The relative gene expression compared with the control plant was calculated using the 2−ΔΔCt method [59].

3. Results

3.1. Phenotypic Data for the DM Analysis

In this study, the DM incidence of DM-resistant genotypes (CML228, Ki3, and Ki11) and DM-susceptible genotypes (B73 and CML270) are reported as phenotypically similar between April and September 2015 (Table 1). Both B73 and CML270 showed 100% DM disease incidence within 6 weeks after inoculation in April and September, while comparable figures for CML228, Ki3, and Ki11 were 0%/25%, 0%/22.2%, and 5%/25%, respectively. According to the weather records, September was wetter and cooler than April.
We evaluated resistance to P. sorghi, P. maydis, and S. macrospora using 192 F7 families derived from B73 (susceptible) × Ki11 (resistant) since we decided that the approach of polygenic control for the various DM strains would offer more effective pathogen resistance (Figure S2).
We performed DM screening of F7 families in September 2017 because DM development is favorable under high humidity and low temperature conditions [60]. The mean of the DM incidence for the F7 families ranged from 0% to 100%, and the distribution of the F7 families was skewed toward the susceptible lines (a RILs mean of 86.49% for DM incidence) (Figure 1 and Table 2).

3.2. Marker Data Analysis and Linkage Mapping

We screened 691 SSR and 36 RFLP markers to identify polymorphisms between B73 and Ki11. A total of 228 SSR and 6 RFLP markers were used to construct a linkage map of F7 families. The map covered around 2042.51 cM at an average marker interval of 9.12 cM and 79 bins among the approximate 120 bin locations on 10 chromosomes. Around 85.04% of the markers were within 20 cM of the nearest interval. Chromosome 2 had the highest number of markers (34), while chromosome 7 had the lowest (14) (Table S5).

3.3. QTL Analysis

QTLs were classified using logarithm of odds (LOD) scores exceeding the 3.0 threshold. The seven QTLs for P. sorghi, P. maydis, and S. macrospora resistance were detected by constructing a linkage map and analyzing the LOD scores for the F7 families (B73 × Ki11) by ICIM-ADD using QTL IciMapping (Figure 2, Figure S3, and Table 3). The seven QTLs were located on chromosomes 2 (bins 2.01 and 2.02), 3 (bins 3.04 and 3.05), 6 (bin 6.05/6.06), and 9 (bins 9.05 and 9.07). All of the QTLs were contributed by the resistant parent (Ki11). The LOD scores and phenotypic variation explained (PVE) values of the QTLs ranged from 3.17 to 18.16 and from 0.47% to 12.95%, respectively. The QTL detected on bin 2.01 had a major effect whereas the ones on bins 2.02, 3.04, 3.05, 6.05/6.06, 9.05, and 9.07 contributed minor effects (the QTLs were classified with PVE values higher than 10.0% as major and less than 10.0% as minor) [49]. The major QTL on qDM1 (2.01) flanked by umc1165 and bnlg1297 presented 12.95% of the phenotypic variation for P. sorghi, P. maydis, and S. macrospora resistance. The other minor QTLs on qDM2 (2.02), qDM3 (3.04), qDM4 (3.05), qDM5 (6.05/6.06), qDM6 (9.05), and qDM7 (9.07) flanked by umc2363-phi098, umc1030-phi243966, mmc0022-bnlg420, bnlg345-umc1859, umc1231-umc2343, and dupssr29-umc1505 presented 0.75%, 0.48%, 2.90%, 2.85%, 0.77%, and 0.47% of the phenotypic variation, respectively.

3.4. qRT-PCR Validation of the Candidate Genes for P. sorghi, P. maydis, and S. macrospora Resistance

We performed an analysis of the expression levels of reference genes (18s rRNA genes, UBCP, MEP, and LUG) [58]. Nineteen DM-related genes from Arabidopsis thaliana, Oryza sativa, and Zea mays (LOC4345959, LOC107275863, SGT1B, LOC103650325, LOC100382073, LOC103647182, LOC4351808, HSK, LOC4324025, LOC103654479, LOC4345309, DMR6, IDC1, LOC103632498, LOC100191339, LOC107275878, LOC103648264, LOC103642860, and EDM2) were validated by comparing control and DM-infected plants that were DM-resistant (CML228, Ki3, and Ki11) or DM-susceptible (B73 and CML270) genotypes (Table 4). Five genes (LOC107275878, LOC103632498, LOC103647182, HSK, and LOC4345309) were upregulated by DM in the DM-resistant genotypes, while another five genes (LOC103648264, LOC4324025, LOC100191339, LOC100382073, and IDC1) were partially upregulated (Figure 3). The other nine genes were not expressed in the control or DM-infected genotypes. We considered that gene expression levels showed a wide range of variation because resistant genotypes (CML228, Ki3, and Ki11) differ in origin and pedigree.
The candidate genes related to P. sorghi, P. maydis, and S. macrospora resistance were obtained by searching for all of the flanked markers in the QTL regions in MaizeGDB (B73 RefGen_v2) (Table 5 and Table S6). The genes were flanked by 14 markers (umc1165, bnlg1297, umc2363, phi098, umc1030, phi243966, mmc0022, bnlg420, bnlg345, umc1859, umc1231, umc2343, dupssr29, and umc1505). Three markers (bnlg1297, umc2363, and phi098) were identified by the same physical location in B73 RefGen_v2, and so we suggest that qDM1 and qDM2 are likely the same. qDM1 was analyzed as a major QTL, thus we supposed that DM-related genes were possible to locate in qDM1 region. qDM1 and qDM2 are located at 4,051,998–4,407,701; 37 genes were obtained in the region. Of the 37 genes, 7 were identified as short sequence length or without genetic information. The relative expression levels of the 62 genes by qRT-PCR were validated by comparing the control and DM-infected plants from B73, CML228, CML270, Ki3, and Ki11. However, 17 genes were not expressed in the control or DM-infected genotypes.
From the 45 genes, 15 (GRMZM2G128315, AC210003.2_FG004, GRMZM2G045049, GRMZM2G178880, GRMZM2G363066, GRMZM2G028643, GRMZM2G330907, GRMZM2G047677, AC191071.3_FG001, GRMZM2G133707, GRMZM2G020043, GRMZM2G039345, GRMZM2G314171, GRMZM2G062031, and GRMZM2G005984) were upregulated by DM in the DM-resistant genotypes (Figure 4 and Figure S4). These 15 upregulated genes were classified as peroxidase (POX), chloroplastic/mitochondrial phosphoribosylformylglycinamidine (PITG) synthase, mannan synthase (ManS), G-type lectin S-receptor-like serine/threonine protein kinase (GsSTK), LRR receptor-like STK (LRR-STK), putative LRR receptor-like protein kinase (LRR-RLK) family protein, LRR family protein, putative STRUBBELIG family receptor protein kinase, abscisic acid receptor PYL5, abscisic acid receptor PYL3, pyrabactin resistance-like protein (PYL), probable flavin-containing monooxygenase 1 (FMOs), P-loop containing nucleoside triphosphate hydrolase superfamily protein (NTPase), photosystem II repair protein PSB27-H1 chloroplastic, photosystem II protein, and uncharacterized. The uncharacterized genes (GRMZM2G062031, GRMZM2G133707, GRMZM2G20043, and GRMZM2G314171) were highly expressed in DM-infected plants. Of the 45 genes, 30 were partially upregulated and were classified into 20 functional annotations.

4. Discussion

DM, caused by Peronosclerospora species in maize, causes severe yield loss despite the use of chemical pesticides and has spread in many tropical and subtropical regions throughout the world. The lack of suitable studies for gene diversity in DM has been a major constraint in tropical Asia, especially in the maize-growing environments of South and Southeast Asia [61]. The results from a great many studies on P. sorghi (sorghum DM) resistance have been reported because it is widely distributed throughout Asia, Africa, and America, where it is easier to perform experiments than in other regions. However, we need to evaluate various DM species because the fungus can easily spread to other regions through the movement of spores in the air.
We screened resistance for P. sorghi (sorghum DM), P. maydis (Java DM), and S. macrospora (crazy top DM) using 192 F7 families derived from B73 (susceptible) × Ki11 (resistant) in Phnom Penh, Cambodia. The distribution of the F7 families was skewed toward the susceptible lines. According to previous studies, it is not uncommon that phenotype values of mapping populations do not follow a normal distribution [2,21,30]. The phenotype data for the DM of the F7 families leaned more toward susceptible because of the response to the three DM pathogens. We used 691 SSR and 36 RFLP markers from MaizeGDB (B73 RefGen_v2) for parental polymorphism. We performed polymorphism analysis by electrophoresis using 3% agarose gel with 1× TAE buffer at 100 V for 1 h. It is necessary that a difference of greater than 20 bp between PCR products could be detected by using 3% agarose gel and loading for 1 h. Additionally, we analyzed flanked markers in the QTL regions by using QiaXcel advanced system (Qiagen, Hilden, Germany). The constructed linkage map covered around 2042.51 cM at average intervals of 9.12 cM between markers for 228 SSR and six RFLP markers on 10 chromosomes. Around one-third of the markers showed to be polymorphic; this result was considered an average ratio when compared to references [21,32,33,34,62]. These markers were distributed over 79 bins among around 120 bin locations. Approximately 85.04% of the markers were within 20 cM of the nearest interval (Table S5). In QTL analysis, it is likely that RILs (F7 families (B73 × Ki11)) made a significant contribution because the use of RILs is more powerful than F2 or backcross populations in QTL analysis [25,26,27]. However, in this study, phenotype data for P. sorghi, P. maydis, and S. macrospora resistance showed recessive trait. In analysis of recessive traits, backcross inbred lines (BILs) derived from resistant genotype have more power to detect recessive QTLs because they have an advantage in segregation ratio compared to F2 or RILs [63].
The seven QTLs for P. sorghi, P. maydis, and S. macrospora resistance were identified on 4 of 10 chromosomes (Table 3). A major QTL was identified in bin 2.01 (qDM1) on chromosome 2, while the other QTLs were detected in bins qDM2 (2.02), qDM3 (3.04), qDM4 (3.05), qDM5 (6.05/6.06), qDM6 (9.05), and qDM7 (9.07). All of the DM-resistant alleles were obtained from Ki11. According to previous studies, the QTLs for DM resistance have been detected in chromosomes 2, 3, 6, and 9, and several flanked markers of QTL regions have been reported. We considered that these seven QTLs affected P. sorghi (sorghum DM), P. maydis (Java DM), and S. macrospora (crazy top DM) from Cambodia. The QTL regions found in the present study were similar to those in previous studies for P. sorghi (sorghum DM) and P. heteropogoni (Rajasthan DM) resistance [21,32,33,34,49] (Table S7). qDM1 and qDM2 shared the same flanked markers (umc1165 and umc2363 in bins 2.01 and 2.02, respectively) [34]. qDM3 had flanked marker umc1030 in bin 3.04 [62]. qDM4 and qDM5 shared the same flanked markers (bnlg420 and umc1859 in bins 3.05 and 6.05/6.06, respectively) [21]. In addition, qDM6 and qDM7 shared the same flanked markers (umc2343 and dupssr29 in bins 9.05 and 9.07, respectively) [33]. The bin locations (3.04 and 3.05, respectively) of qDM3 and qDM4 matched with other reports on QTLs [21,32,33,49]. The other QTLs were identified from five different DM strains (P. sorghi and P. heteropogoni from India, P. zeae from Thailand, P. philippinensis from the Philippines, and P. maydis from Indonesia) and sorghum DM on bin 6.05 using RILs (derived from Ki3) and backcross populations [2,21]. Ki3 and Ki11 were developed as DM-resistant lines derived from the Suwan1 strain from Thailand. These results suggest that it should be possible to detect candidate genes for P. sorghi, P. maydis, and S. macrospora resistance near these bin positions. Although the QTL analysis was performed using different mapping populations, the results of this study are well matched with previously reported ones.
We analyzed 19 DM-related genes from Arabidopsis thaliana, Oryza sativa, and Zea mays (Table 4): EDM2, SGT1B, LOC100191339, HSK, DMR6, IDC1, and LOC103642860, among others. Among the 19 genes, five were upregulated in the DM-resistant genotypes, two of which (LOC103632498 and LOC103647182) are located on chromosomes 2 and 7 of maize, respectively. LOC100191339 was located on nearby qDM4 and was highly upregulated in DM-infected CML270 (susceptible). Two genes (LOC107275878 and LOC4345309) and HSK, originating from Arabidopsis thaliana and Oryza sativa, were significantly upregulated in the DM-infected DM-resistant genotypes. Hence, we considered that these genes are conserved in monocotyledon and dicotyledon plants [64,65].
We obtained 62 candidate genes for P. sorghi, P. maydis, and S. macrospora resistance near the flanked markers in the QTL region; these genes were validated by comparing their relative expression levels in the control and DM-infected groups. The physical locations of candidate genes were continually updated from B73 RefGen_v1 to B73 RefGen_v4, but updating the genetic information of the genomic markers is slower than the transcripts. Hence, we used B73 RefGen_v2 to set the physical locations of the genomic markers and candidate genes. The annotations of 45 genes were analyzed to predict their functions using the Pfam database. The annotations and functions of these genes were identified as being related to DM resistance (Table 6 and Table S8). There are various factors involved in DM resistance, such as POX (peroxidase), MYB (transcriptional activator Myb), GSO1 (LRR receptor-like serine/threonine protein kinases (STKs), (LRR-STKs)), plant RLKs, polygalacturonase inhibitor protein (PGIP), polyphenol oxidase (PPO), NAC, WRKY transcription factors, and pathogenesis-related (PR) protein [66,67,68,69,70,71,72].
Three genes (AC210003.2_FG004, AC191071.3_FG001, and GRMZM2G039345) were annotated as POX, FMO, and RuBisCO (ribulose-1,5-bisphosphate carboxylase/oxygenase), respectively. The activities of POX and PPO, along with b-1,3-glucanase, are associated with DM resistance in sunflowers [73]. Plants can implement DM post-infection mechanisms such as an increase in localized callose deposition to fortify plant cell walls [74,75], reactive oxygen species (ROS), peroxidase activity, and hypersensitive response activation [74,76]. When a resistant grapevine is infected with DM, it produces high concentrations of resveratrol that can be oxidized by induced peroxidase [77]. In broccoli, BoAPX (ascorbate peroxidase) genes contribute enhanced both DM and heat tolerance, and play important roles in cellular defense against ROS-mediate oxidative damage [68]. In previous studies, high POX activity has been associated with resistance to PM [78] in lettuce [79] and melon [80] and to Verticillium dahlia in tomatoes [81]. PM is very similar to DM in that both are caused by fungi and are widespread under humid and low temperature conditions; the disease symptoms are also similar. FMO1 positively regulates the enhanced disease susceptibility1 (EDS1) pathway in Arabidopsis thaliana. The EDS1 pathway controls defense activation and programmed cell death against pathogens [82]. In addition, a defect in FMO1 partially disables toll interleukin 1 receptor nucleotide binding sites leucine-rich repeat (TIR-NB-LRR) resistance and basal defense. In Arabidopsis thaliana, RPP4, which has been identified as a TIR-NB-LRR protein coupled with its dependence on signaling components in leaves, confers resistance to DM [83]. RPP4-mediated resistance is regulated by interaction between EDS1 and NDR1 signaling in cotyledons.
Two genes (GRMZM2G342564 and GRMZM2G040095) were identified as lipoxygenase (LOX)-producing. LOX is known to play a role in disease resistance for many host pathosystems. In earlier reports, LOX activity was found to increase in resistant plants and to decrease in susceptible plants. Also, LOX was reported to affect DM and PM resistance of pearl millet and wheat, respectively [84,85,86,87,88].
GRMZM2G028643, GRMZM2G128315, and GRMZM2G330907 (identified as LRR) and GRMZM2G363066 (identified as nonspecific STK) were associated with defense reactions against pathogens. PR protein are encoded by disease resistance (R) genes, responding to pathogenic microorganisms and signaling cascades that activate defense reactions [89,90]. The largest family of PR proteins is defined by the presence of 12 to 21 LRRs (it has been speculated that LRRs bind pathogen-derived ligands). A glutamate-to-lysine substitution in LRR partially compromises the function of R genes against DM [91]. Several kinases (protein kinases, wall-associated kinases, calcium-dependent protein kinases, STKs, LRR-STKs, and mitogen-activated protein kinases (MAPKs)) are strongly associated with signal transduction mediated by Ca2+ permeable channels [90,92,93,94]. It has been shown that the activation of the PR protein and phenylpropanoid pathway enzymes such as LRR-STKs and MAPKs responds to DM infection in pearl millet by inducing molecules for signal transduction [95]. Also, nucleotide-binding site (NBS)-LRR and receptor-like proteins (RLP) were included in five classes of R genes, these have been reported as resistance against DM [70,71]. RLKs are well known to play a role in many important signaling process such as plant growth, development, hormone signaling, and stress response. LRR-RLK family proteins regulate plant innate immunity and defense [69]. PGIP is a defense protein consisting of an extra-cytoplasmic LRR which specifically binds the invading fungus cell wall to the host tissue [96]. In previous studies, the transcription of PGIP and a cell wall glycoprotein were induced in DM-resistant strains of pearl millet [38] and grape [97]. In Arabidopsis thaliana, homoserine accumulation in the chloroplast triggers a novel form of DM resistance that is independent of known immune responses [98]. In pearl millet, the role of PGIP in resistance against DM pathogen (Sclerospora graminicola) was reported by differential gene expression analysis between resistant and susceptible genotypes [72].
GRMZM2G005984 was classified as a photosystem II protein. In grapevine, PM-responsive proteins are involved in photosynthesis, metabolism, disease/defense, protein destination, and protein synthesis [99]. These proteins are associated with the plant defense response and slow down disease progression against Erysiphe necator. GRMZM2G314171 was classified as DEAD-box RNA helicase, one of which (OsBIRH1) has been found to modulate the defensive response to infection and oxidative stress in rice [100]. The DEAD-box RNA helicase family functions in chloroplast biogenesis in maize [101]. GRMZM2G178880 was classified as a GTF (general transcription factor); these proteins are probably involved in defense-related processes [102]. It has been shown that GTF expression increases during the interaction between PM and barley [103], and its expression level is induced after wounding and phytopathogenic bacteria attack [104]. GRMZM2G024293 was identified as a conserved hypothetical adenosine triphosphate (ATP)-binding protein. Its downregulation brings about a decrease in ATP binding in DM-infected plants, which makes it possible to estimate DM susceptibility. Meanwhile, further study is needed to identify the functional characteristics of the uncharacterized genes.
Hence, we suggest that 10 genes (AC210003.2_FG004, AC191071.3_FG001, GRMZM2G039345, GRMZM2G028643, GRMZM2G128315, GRMZM2G330907, GRMZM2G363066, GRMZM2G005984, GRMZM2G178880, and GRMZM2G314171) are related to P. sorghi, P. maydis, and S. macrospora resistance. The downregulation of some of the P. sorghi, P. maydis, and S. macrospora resistance genes occurs in DM-infected plants. Moreover, it was shown that 30 genes are partially upregulated in DM-infected plants of resistant genotype. From these results, we predict the possibility that the phenotype of resistant plants shows resistance (R) or moderate resistance (MR).
Experiments on P. sorghi, P. maydis, and S. macrospora resistance have been limited to QTL analysis in previous studies. However, we focused on their great importance through the screening of candidate genes using qRT-PCR. The results of this study can serve as fine-mapping for DM and marker-assisted selection (MAS) in maize breeding. In further research, our approach could be used to screen other DM strains in different environments using RILs and to validate candidate genes using CRISPR/Cas9. In addition, the genes and pathways associated with resistance could be identified via comparative transcriptome profiling using RNA-seq.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4425/11/2/191/s1, Figure S1: Field layout using the spreader-row technique in Phnom Penh, Cambodia; Figure S2: Maize phenotypes and symptoms of downy mildew (DM) 6 weeks after inoculation using the spreader-row technique; Figure S3: The logarithm of odds (LOD) values of the P. sorghi, P. maydis, and S. macrospora resistance quantitative trail loci (QTLs) between the flanked markers; Figure S4: Relative expression levels (log2DM-infected/control) of 30 candidate genes for P. sorghi, P. maydis, and S. macrospora via quantitative real-time polymerase chain reaction (qRT-PCR) on the control and infected plants; Table S1: Primer information of SSR and RFLP markers for QTL analysis; Table S2: Primer information of P. sorghi, P. maydis, and S. macrospora related genes for qRT-PCR analysis; Table S3: Primer information of candidate genes for qRT-PCR analysis; Table S4: Primer information of reference genes for qRT-PCR analysis; Table S5: The numbers of markers and bins covered in each chromosome; Table S6: List of the P. sorghi, P. maydis, and S. macrospora resistance candidate genes between the left and right flanked markers in QTL regions; Table S7: Comparison of QTLs identified in the present study with previous reports for resistance to P. sorghi, P. maydis, and S. macrospora; Table S8: Pfam (EBI 2019) domain analysis of 47 candidate genes.

Author Contributions

H.C.K. and K.-H.K. designed and performed the field experiments and sampling. H.C.K. conducted the QTL and gene analysis work and wrote the manuscript. K.S., and J.Y.K. assisted with the QTL and gene analysis work. B.-M.L. advised on the experiments and data analysis. All authors have read and agreed to the published version of the manuscript.

Funding

This work was carried out with the support of the Cooperative Research Program for Agriculture Science & Technology Development (Project No. PJ012649022020), Rural Development Administration, Republic of Korea, and was also supported by the Golden Seed Project (213009-05-4-SB710); Ministry of Agriculture, Food, and Rural Affairs (MAFRA); Ministry of Oceans and Fisheries (MOF); Rural Development Administration (RDA); and Korea Forest Services (KFS) in the Republic of Korea.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bonde, M. Epidemiology of downy mildew diseases of maize, sorghum and pearl millet. Trop. Pest Manag. 1982, 28, 49–60. [Google Scholar] [CrossRef]
  2. George, M.; Prasanna, B.; Rathore, R.; Setty, T.; Kasim, F.; Azrai, M.; Vasal, S.; Balla, O.; Hautea, D.; Canama, A. Identification of QTLs conferring resistance to downy mildews of maize in Asia. Theor. Appl. Genet. 2003, 107, 544–551. [Google Scholar] [CrossRef] [PubMed]
  3. Jeffers, D.; Cordova, H.; Vasal, S.; Srinivasan, G.; Beck, D.; Barandiaran, M. Status in breeding for resistance to maize diseases at CIMMYT. In Proceedings of the 7th Asian Regional Maize Workshop, Los Baños, Phillipines, 23–27 February 1998; Vasal, S.K., Gonzalez, C.F., Fan, X.M., Eds.; CIMMYT: El Batan, Mexico, 2000; pp. 257–266. [Google Scholar]
  4. Centro Internacional de Mejoramiento de Maíz y Trigo (CIMMYT); Pingali, P.L. CIMMYT 1999/2000 World Maize Facts and Trends-Meeting World Maize Needs: Technological Opportunities and Priorities for the Public Sector; CIMMYT: El Batan, Mexico, 2001. [Google Scholar]
  5. Rathore, R.; Siradhana, B. Estimation of losses caused by Peronosclerospora heteropogoni on Ganga-5 maize hybrid. Phytophylactica 1988, 19, 119–120. [Google Scholar]
  6. Rifin, A. Downy mildew resistance of single cross progenies between Indonesia and Philippine corn inbred lines. Penelitian Pertanian 1983, 3, 81–83. [Google Scholar]
  7. Broyles, J. Observations on time and location of penetration in relation to amount of damage and chemical control of Physoderma maydis. Phytopathology 1956, 46, 8. [Google Scholar]
  8. Borges, F.; Orangel, L. Diallel analysis of maize resistance to sorghum downy mildew 1. Crop Sci. 1987, 27, 178–180. [Google Scholar] [CrossRef]
  9. Frederiksen, R.; Renfro, B. Global status of maize downy mildew. Ann. Rev. Phytopathol. 1977, 15, 249–271. [Google Scholar] [CrossRef]
  10. Payak, M.; Renfro, B. A new downy mildew disease of maize. Phytopathology 1967, 57, 394–397. [Google Scholar]
  11. Sharma, R.; De Leon, C.; Payak, M. Diseases of maize in South and South-East Asia: Problems and progress. Crop Prot. 1993, 12, 414–422. [Google Scholar] [CrossRef]
  12. Spencer, M.; Dick, M. Aspects of graminicolous downy mildew biology: Perspectives for tropical plant pathology and Peronosporomycetes phylogeny. Trop. Mycol. 2002, 2, 63–81. [Google Scholar]
  13. Yen, T.T.O.; Prasanna, B.; Setty, T.; Rathore, R. Genetic variability for resistance to sorghum downy mildew (Peronosclerospora sorghi) and Rajasthan downy mildew (P. heteropogoni) in the tropical/sub-tropical Asian maize germplasm. Euphytica 2004, 138, 23–31. [Google Scholar] [CrossRef]
  14. Ullstrup, A. Observations on crazy top of Corn. Phytopathology 1952, 42, 675–680. [Google Scholar]
  15. Putnam, M.L. Brown stripe downy mildew (Sclerophthora rayssiae var. zeae) of maize. Plant Health Prog. 2007, 8, 34. [Google Scholar] [CrossRef]
  16. Geetha, K.; Jayaraman, N. Inheritance of sorghum downy mildew resistance in maize. Indian J. Agric. Res. 2002, 36, 234–240. [Google Scholar]
  17. Jinahyon, S. The genetics of resistance and its implications to breeding for resistance in corn. In Proceedings of the Inter-Asian Corn Improvement Workshop, Kuala Lampur, Malyasia, 10–12 December 1973; pp. 30–39. [Google Scholar]
  18. Kaneko, K.; Aday, B. Inheritance of resistance to philippine downy mildew of maize, Peronosclerospora philippinensis 1. Crop Sci. 1980, 20, 590–594. [Google Scholar] [CrossRef]
  19. Leon, C.D.; Ahuja, V.; Capio, E.R.; Mukherjee, B. Genetics of resistance to Philippine downy mildew in three maize populations. Indian J. Gen. Plant Breeding 1993, 53, 406–410. [Google Scholar]
  20. Nair, S.K.; Prasanna, B.; Rathore, R.; Setty, T.; Kumar, R.; Singh, N. Genetic analysis of resistance to sorghum downy mildew and Rajasthan downy mildew in maize (Zea mays L.). Field Crops Res. 2004, 89, 379–387. [Google Scholar] [CrossRef]
  21. Nair, S.K.; Prasanna, B.M.; Garg, A.; Rathore, R.; Setty, T.; Singh, N. Identification and validation of QTLs conferring resistance to sorghum downy mildew (Peronosclerospora sorghi) and Rajasthan downy mildew (P. heteropogoni) in maize. Theor. Appl. Genet. 2005, 110, 1384–1392. [Google Scholar] [CrossRef]
  22. Nallathambi, P.; Sundaram, K.M.; Arumugachamy, S. Inheritance of Resistance to Sorghum Downy mildew (Peronosclerospera sorghi) in Maize (Zea mays L.). Int. J. Agric. Environ. Biotechnol. 2010, 3, 285–293. [Google Scholar]
  23. Singburaudom, N.; Renfro, B. Heritability of resistance in maize to sorghum downy mildew (Peronosclerospora sorghi (Weston and Uppal) CG Shaw). Crop Prot. 1982, 1, 323–332. [Google Scholar] [CrossRef]
  24. Young, N. QTL mapping and quantitative disease resistance in plants. Ann. Rev. Phytopathol. 1996, 34, 479–501. [Google Scholar] [CrossRef] [PubMed]
  25. Austin, D.; Lee, M. Comparative mapping in F 2:3 and F 6:7 generations of quantitative trait loci for grain yield and yield components in maize. Theor. Appl. Genet. 1996, 92, 817–826. [Google Scholar] [CrossRef] [PubMed]
  26. Burr, B.; Burr, F.A. Recombinant inbreds for molecular mapping in maize: Theoretical and practical considerations. Trends Genet. 1991, 7, 55–60. [Google Scholar] [CrossRef]
  27. Knapp, S.; Bridges, W. Using molecular markers to estimate quantitative trait locus parameters: Power and genetic variances for unreplicated and replicated progeny. Genetics 1990, 126, 769–777. [Google Scholar] [PubMed]
  28. Krakowsky, M.; Lee, M.; Woodman-Clikeman, W.; Long, M.; Sharopova, N. QTL mapping of resistance to stalk tunneling by the European corn borer in RILs of maize population B73× De8 1. Crop Sci. 2004, 44, 274–282. [Google Scholar]
  29. Frova, C.; Gorla, M.S. Quantitative expression of maize HSPs: Genetic dissection and association with thermotolerance. Theor. Appl. Genet. 1993, 86, 213–220. [Google Scholar] [CrossRef]
  30. Guo, J.; Su, G.; Zhang, J.; Wang, G. Genetic analysis and QTL mapping of maize yield and associate agronomic traits under semi-arid land condition. Afr. J. Biotechnol. 2008, 7, 1829–1838. [Google Scholar]
  31. Agrama, H.; Moussa, M.; Naser, M.; Tarek, M.; Ibrahim, A. Mapping of QTL for downy mildew resistance in maize. Theor. Appl. Genet. 1999, 99, 519–523. [Google Scholar] [CrossRef]
  32. Sabry, A.; Jeffers, D.; Vasal, S.; Frederiksen, R.; Magill, C. A region of maize chromosome 2 affects response to downy mildew pathogens. Theor. Appl. Genet. 2006, 113, 321–330. [Google Scholar] [CrossRef]
  33. Jampatong, C.; Jampatong, S.; Jompuk, C.; Sreewongchai, T.; Grudloyma, P.; Balla, C.; Prodmatee, N. Mapping of QTL affecting resistance against sorghum downy mildew (Peronosclerospora sorghi) in maize (Zea mays L). Maydica 2013, 58, 119–126. [Google Scholar]
  34. Lohithaswa, H.; Jyothi, K.; Kumar, K.S.; Hittalmani, S. Identification and introgression of QTLs implicated in resistance to sorghum downy mildew (Peronosclerospora sorghi (Weston and Uppal) CG Shaw) in maize through marker-assisted selection. J. Genet. 2015, 94, 741–748. [Google Scholar] [CrossRef] [PubMed]
  35. Li, J.; Ding, Q.; Wang, F.; Li, H.; Zhang, Y.; Liu, L.; Jiao, Z.; Gao, J. Genome-wide gene expression profiles in response to downy mildew in Chinese cabbage (Brassica rapa L. ssp. pekinensis). Eur. J. Plant Pathol. 2018, 151, 861–873. [Google Scholar] [CrossRef]
  36. Burkhardt, A.; Day, B. Transcriptome and small RNAome dynamics during a resistant and susceptible interaction between cucumber and downy mildew. Plant Genome 2016, 9, 1–19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Li, X.; Wu, J.; Yin, L.; Zhang, Y.; Qu, J.; Lu, J. Comparative transcriptome analysis reveals defense-related genes and pathways against downy mildew in Vitis amurensis grapevine. Plant Physiol. Biochem. 2015, 95, 1–14. [Google Scholar] [CrossRef]
  38. Kulkarni, K.S.; Zala, H.N.; Bosamia, T.C.; Shukla, Y.M.; Kumar, S.; Fougat, R.S.; Patel, M.S.; Narayanan, S.; Joshi, C.G. De novo transcriptome sequencing to dissect candidate genes associated with pearl millet-downy mildew (Sclerospora graminicola Sacc.) interaction. Front. Plant Sci. 2016, 7, 847. [Google Scholar] [CrossRef] [Green Version]
  39. Okubara, P.A.; Anderson, P.A.; Ochoa, O.E.; Michelmore, R.W. Mutants of downy mildew resistance in Lactuca sativa (lettuce). Genetics 1994, 137, 867–874. [Google Scholar]
  40. Zhu, Q.; Gao, P.; Wan, Y.; Cui, H.; Fan, C.; Liu, S.; Luan, F. Comparative transcriptome profiling of genes and pathways related to resistance against powdery mildew in two contrasting melon genotypes. Sci. Hortic. 2018, 227, 169–180. [Google Scholar] [CrossRef]
  41. Aekatasanawan, C.; Jampatong, S.; Aekatasanawan, C.; Chulchoho, N.; Balla, C. Supporting the hybrid maize breeding research in Thailand. In Proceedings of the 7th Asian Regional Maize Workshop, Los Baños, Phillipines, 23–27 February 1998; Vasal, S.K., Gonzalez, C.F., Fan, X.M., Eds.; CIMMYT: El Batan, Mexico, 2000; pp. 82–91. [Google Scholar]
  42. Hallauer, A. Specialty Corns, 2nd ed.; CRC Press: Boca Raton, FL, USA, 2001. [Google Scholar]
  43. Coles, N.D.; McMullen, M.D.; Balint-Kurti, P.J.; Pratt, R.C.; Holland, J.B. Genetic control of photoperiod sensitivity in maize revealed by joint multiple population analysis. Genetics 2010, 184, 799–812. [Google Scholar] [CrossRef] [Green Version]
  44. Kim, K.; Moon, J.; Kim, J.; Kim, H.; Shin, S.; Song, K.; Baek, S.; Lee, B. Evaluation of maize downy mildew using spreader row technique. Korean J. Crop Sci. 2016, 61, 41–49. [Google Scholar] [CrossRef] [Green Version]
  45. Liu, K.; Goodman, M.; Muse, S.; Smith, J.S.; Buckler, E.; Doebley, J. Genetic structure and diversity among maize inbred lines as inferred from DNA microsatellites. Genetics 2003, 165, 2117–2128. [Google Scholar]
  46. Russell, W. Registration of B70 and B73 parental lines of Maize1 (Reg. Nos. PL16 and PL17). Crop Sci. 1972, 12, 721. [Google Scholar] [CrossRef]
  47. Kim, J.Y.; Moon, J.C.; Kim, H.C.; Shin, S.; Song, K.; Kim, K.H.; Lee, B.M. Identification of downy mildew resistance gene candidates by positional cloning in maize (Zea mays subsp. mays; Poaceae). Appl. Plant Sci. 2017, 5, 1600132. [Google Scholar] [CrossRef] [PubMed]
  48. Craig, A.; Bockholt, J.; Frederiksen, R.A.; Zuber, M.S. Reaction of important corn inbred lines to Peronosclerospora sorghi. Plant Dis. Rept. 1977, 61, 563–564. [Google Scholar]
  49. Nagabhushan, T.; Lohithaswa, H.; Sreeramasetty, T.; Puttaramanaik, H.S.; Hittalmani, S. Identification of stable source of resistance to sorghum downy mildew in maize (Zea mays L.). J. Agroecol. Nat. Resour. Manag. 2014, 1, 176–178. [Google Scholar]
  50. Rao, B.; Prakash, H.; Shetty, H. Relationship of cultivars with sporulation and morphology of asexual propagules of Peronosclerospora sorghi on maize. Int. J. Trop. Plant Dis. 1984, 2, 175–180. [Google Scholar]
  51. Li, H.; Ye, G.; Wang, J. A modified algorithm for the improvement of composite interval mapping. Genetics 2007, 175, 361–374. [Google Scholar] [CrossRef] [Green Version]
  52. Wang, J.; Li, H.; Zhang, L.; Meng, L. User’s Manual of QTL IciMapping ver. 4.1; CIMMYT: El Batan, Mexico, 2016. [Google Scholar]
  53. Kosambi, D.D. The estimation of map distances from recombination values. Ann. Eugen. 1944, 12, 172–175. [Google Scholar] [CrossRef]
  54. Doerge, R.W.; Churchill, G.A. Permutation tests for multiple loci affecting a quantitative character. Genetics 1996, 142, 285–294. [Google Scholar]
  55. Cui, F.; Fan, X.; Chen, M.; Zhang, N.; Zhao, C.; Zhang, W.; Han, J.; Ji, J.; Zhao, X.; Yang, L. QTL detection for wheat kernel size and quality and the responses of these traits to low nitrogen stress. Theor. Appl. Genet. 2016, 129, 469–484. [Google Scholar] [CrossRef]
  56. Lander, E.; Kruglyak, L. Genetic dissection of complex traits: Guidelines for interpreting and reporting linkage results. Nat. Genet. 1995, 11, 241. [Google Scholar] [CrossRef]
  57. Rozen, S.; Skaletsky, H. Primer3 on the WWW for general users and for biologist programmers. In Bioinformatics Methods and Protocols; Springer: Berlin/Heidelberg, Germany, 2000; pp. 365–386. [Google Scholar]
  58. Manoli, A.; Sturaro, A.; Trevisan, S.; Quaggiotti, S.; Nonis, A. Evaluation of candidate reference genes for qPCR in maize. J. Plant Physiol. 2012, 169, 807–815. [Google Scholar] [CrossRef] [PubMed]
  59. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2− ΔΔCT method. Methods 2001, 25, 402–408. [Google Scholar] [CrossRef] [PubMed]
  60. Shekhar, M.; Kumar, S. Inoculation Methods and Disease Rating Scales for Maize Diseases; Directorate of Maize Research: New Delhi, India, 2012. [Google Scholar]
  61. Rashid, Z.; Singh, P.K.; Vemuri, H.; Zaidi, P.H.; Prasanna, B.M.; Nair, S.K. Genome-wide association study in Asia-adapted tropical maize reveals novel and explored genomic regions for sorghum downy mildew resistance. Sci. Rep. 2018, 8, 366. [Google Scholar] [CrossRef] [PubMed]
  62. Nagabhushan; Lohithaswa, H.C.; Pandravada, A.S. Construction of high-density linkage map and identification of QTLs for resistance to sorghum downy mildew in maize (Zea mays L.). Mol. Breeding 2017, 37, 2. [Google Scholar] [CrossRef]
  63. Jeuken, M.J.W.; Pelgrom, K.; Stam, P.; Lindhout, P. Efficient QTL detection for nonhost resistance in wild lettuce. Theor. Appl. Genet. 2008, 116, 845–857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Magalhaes, J.V. Aluminum tolerance genes are conserved between monocots and dicots. Proc. Natl. Acad. Sci. USA 2006, 103, 9749–9750. [Google Scholar] [CrossRef] [Green Version]
  65. Miller, R.N.; Bertioli, D.J.; Baurens, F.C.; Santos, C.M.; Alves, P.C.; Martins, N.F.; Togawa, R.C.; Júnior, M.T.S.; Júnior, G.J.P. Analysis of non-TIR NBS-LRR resistance gene analogs in Musa acuminata Colla: Isolation, RFLP marker development, and physical mapping. BMC Plant Biol. 2008, 8, 15. [Google Scholar] [CrossRef] [Green Version]
  66. Körösi, K.; Bán, R.; Barna, B.; Virányi, F. Biochemical and molecular changes in downy mildew-infected sunflower triggered by resistance inducers. J. Phytopathol. 2011, 159, 471–478. [Google Scholar] [CrossRef]
  67. Toffolatti, S.L.; De Lorenzis, G.; Costa, A.; Maddalena, G.; Passera, A.; Bonca, M.C.; Pindo, M.; Stefani, E.; Cestaro, A.; Casati, P.; et al. Uique resistance traits against downy mildew from the center of origin of grapevine (Vitis vinifera). Sci. Rep. 2018, 8, 1–11. [Google Scholar] [CrossRef]
  68. Jiang, M.; Jiang, J.J.; He, C.M.; Guan, M. Broccoli plants over-expressiong a cytosolic ascorbate peroxidase gene increase resistance to downy mildew and heat stress. J. Plant Pathol. 2016, 98, 413–420. [Google Scholar]
  69. Lin, H.; Leng, H.; Guo, Y.; Kondo, S.; Zhao, Y.; Guo, X. QTLs and candidate genes for downy mildew resistance conferred by interspecific grape (V. vinifera L. x V. amurensis Rupr.) crossing. Sci. Hortic. 2019, 244, 200–207. [Google Scholar] [CrossRef]
  70. She, H.; Quan, W.; Zhang, H.; Liu, Z.; Wang, X.; Wu, J.; Feng, C.; Correll, J.C.; Xu, Z. Fine mapping and candidate gene screening of the downy mildew resistance gene RPF1 in spinach. Theor. Appl. Genet. 2018, 131, 2529–2541. [Google Scholar] [CrossRef] [PubMed]
  71. Zhang, S.; Ding, F.; Peng, H.; Huang, Y.; Lu, J. Molecular cloning of a CC-NBS-LRR gene from Vitis quinquangularis and its expression pattern in response to downy mildew pathogen infection. Mol. Genet. Genomics 2017, 293, 61–68. [Google Scholar] [CrossRef] [PubMed]
  72. Prabhu, S.A.; Kini, K.R.; Raj, S.N.; Moerschbacher, B.M.; Shetty, H.S. Polygalacturonase-inhibitor proteins in pearl millet: Possible involvement in resistance against downy mildew. Acta Biochim. Biophys. Sin. 2012, 44, 415–423. [Google Scholar] [CrossRef] [Green Version]
  73. Nandeeshkumar, P.; Sudisha, J.; Ramachandra, K.K.; Prakash, H.; Niranjana, S.; Shekar, S.H. Chitosan induced resistance to downy mildew in sunflower caused by Plasmopara halstedii. Physiol. Mol. Plant Pathol. 2008, 72, 188–194. [Google Scholar] [CrossRef]
  74. Diez-Navajas, A.; Wiedemann-Merdinoglu, S.; Greif, C.; Merdinoglu, D. Nonhost versus host resistance to the grapevine downy mildew, Plasmopara viticola, studied at the tissue level. Phytopathology 2008, 98, 776–780. [Google Scholar] [CrossRef] [Green Version]
  75. Jürges, G.; Kassemeyer, H.H.; Dürrenberger, M.; Düggelin, M.; Nick, P. The mode of interaction between Vitis and Plasmopara viticola Berk. & Curt. Ex de Bary depends on the host species. Plant Biol. 2009, 11, 886–898. [Google Scholar]
  76. Kortekamp, A. Expression analysis of defence-related genes in grapevine leaves after inoculation with a host and a non-host pathogen. Plant Physiol. Biochem. 2006, 44, 58–67. [Google Scholar] [CrossRef]
  77. Pezet, R.; Gindro, K.; Viret, O.; Spring, J.-L. Glycosylation and oxidative dimerization of resveratrol are respectively associated to sensitivity and resistance of grapevine cultivars to downy mildew. Physiol. Mol. Plant Pathol. 2004, 65, 297–303. [Google Scholar] [CrossRef]
  78. Reuveni, R.; Bothma, G. The relationship between peroxidase activity and resistance to Sphaerotheca fuliginea in melons. J. Phytopathol. 1985, 114, 260–267. [Google Scholar] [CrossRef]
  79. Reuveni, R.; Shimoni, M.; Crute, I. An association between high peroxidase activity in lettuce (Lactuca sativa) and field resistance to downy mildew (Bremia lactucae). J. Phytopathol. 1991, 132, 312–318. [Google Scholar] [CrossRef]
  80. Reuveni, R.; Karchi, Z. Peroxidase activity-a possible marker for resistance of melon against downy mildew. Phytopathology 1987, 77, 1724. [Google Scholar]
  81. Reuveni, R.; Ferreira, J. The relationship between peroxidase activity and the resistance of tomatoes (Lycopersicum esculentum) to Verticillium dahliae. J. Phytopathol. 1985, 112, 193–197. [Google Scholar] [CrossRef]
  82. Bartsch, M.; Gobbato, E.; Bednarek, P.; Debey, S.; Schultze, J.L.; Bautor, J.; Parker, J.E. Salicylic acid–independent enhanced disease susceptibility1 signaling in Arabidopsis immunity and cell death is regulated by the monooxygenase FMO1 and the nudix hydrolase NUDT7. Plant Cell 2006, 18, 1038–1051. [Google Scholar] [CrossRef] [Green Version]
  83. Van Der Biezen, E.A.; Freddie, C.T.; Kahn, K.; Parker, J.E.; Jones, J.D. Arabidopsis RPP4 is a member of the RPP5 multigene family of TIR-NB-LRR genes and confers downy mildew resistance through multiple signalling components. Plant J. 2002, 29, 439–451. [Google Scholar] [CrossRef] [Green Version]
  84. Babitha, M.; Prakash, H.; Shetty, H. Purification and properties of lipoxygenase induced in downy mildew resistant pearl millet seedlings due to infection with Sclerospora graminicola. Plant Sci. 2004, 166, 31–39. [Google Scholar] [CrossRef]
  85. Nagatathna, K.C.; Shetty, S.A.; Bhat, S.G.; Shetty, H.S. The possible involvement of lipoxygenase in downy mildew resistance in pearl millet. J. Exp. Bot. 1992, 43, 1283–1287. [Google Scholar] [CrossRef]
  86. Pushpalatha, H.; Sudisha, J.; Geetha, N.; Amruthesh, K.; Shetty, H.S. Thiamine seed treatment enhances LOX expression, promotes growth and induces downy mildew disease resistance in pearl millet. Biol. Plantarum 2011, 55, 522–527. [Google Scholar] [CrossRef]
  87. Shivakumar, P.; Geetha, H.; Shetty, H. Peroxidase activity and isozyme analysis of pearl millet seedlings and their implications in downy mildew disease resistance. Plant Sci. 2003, 164, 85–93. [Google Scholar] [CrossRef]
  88. Mustafa, G.; Khong, N.G.; Tisserant, B.; Randoux, B.; Fontaine, J.; Magnin-Robert, M.; Reignault, P.; Sahraoui, A.L.H. Defence mechanisms associated with mycorrhiza-induced resistance in wheat against powdery mildew. Funct. Plant Biol. 2017, 44, 443–454. [Google Scholar] [CrossRef] [Green Version]
  89. Dangl, J.L.; Jones, J.D. Plant pathogens and integrated defence responses to infection. Nature 2001, 411, 826. [Google Scholar] [CrossRef] [PubMed]
  90. Hammond-Kosack, K.E.; Parker, J.E. Deciphering plant–pathogen communication: Fresh perspectives for molecular resistance breeding. Curr. Opin. Biotechnol. 2003, 14, 177–193. [Google Scholar] [CrossRef]
  91. Warren, R.F.; Henk, A.; Mowery, P.; Holub, E.; Innes, R.W. A mutation within the leucine-rich repeat domain of the Arabidopsis disease resistance gene RPS5 partially suppresses multiple bacterial and downy mildew resistance genes. Plant Cell 1998, 10, 1439–1452. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Heath, M.C. Nonhost resistance and nonspecific plant defenses. Curr. Opin. Plant Biol. 2000, 3, 315–319. [Google Scholar] [CrossRef]
  93. King, S.R.; McLellan, H.; Boevink, P.C.; Armstrong, M.R.; Bukharova, T.; Sukarta, O.; Win, J.; Kamoun, S.; Birch, P.R.; Banfield, M.J. Phytophthora infestans RXLR effector PexRD2 interacts with host MAPKKKε to suppress plant immune signaling. Plant Cell 2014, 26, 1345–1359. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. White, P.; Broadley, M. Calcium in plants. Ann. Bot. 2003, 92, 487–511. [Google Scholar] [CrossRef] [PubMed]
  95. Melvin, P.; Prabhu, S.A.; Anup, C.P.; Shailasree, S.; Shetty, H.S.; Kini, K.R. Involvement of mitogen-activated protein kinase signalling in pearl millet–downy mildew interaction. Plant Sci. 2014, 214, 29–37. [Google Scholar] [CrossRef]
  96. Di Matteo, A.; Bonivento, D.; Tsernoglou, D.; Federici, L.; Cervone, F. Polygalacturonase-inhibiting protein (PGIP) in plant defence: A structural view. Phytochemistry 2006, 67, 528–533. [Google Scholar] [CrossRef]
  97. Wu, J.; Zhang, Y.; Zhang, H.; Huang, H.; Folta, K.M.; Lu, J. Whole genome wide expression profiles of Vitis amurensis grape responding to downy mildew by using Solexa sequencing technology. BMC Plant Biol. 2010, 10, 234. [Google Scholar] [CrossRef] [Green Version]
  98. Van Damme, M.; Zeilmaker, T.; Elberse, J.; Andel, A.; de Sain-van der Velden, M.; Van den Ackerveken, G. Downy mildew resistance in Arabidopsis by mutation of homoserine kinase. Plant Cell 2009, 21, 2179–2189. [Google Scholar] [CrossRef] [Green Version]
  99. Marsh, E.; Alvarez, S.; Hicks, L.M.; Barbazuk, W.B.; Qiu, W.; Kovacs, L.; Schachtman, D. Changes in protein abundance during powdery mildew infection of leaf tissues of Cabernet Sauvignon grapevine (Vitis vinifera L.). Proteomics 2010, 10, 2057–2064. [Google Scholar] [CrossRef] [PubMed]
  100. Li, D.; Liu, H.; Zhang, H.; Wang, X.; Song, F. OsBIRH1, a DEAD-box RNA helicase with functions in modulating defence responses against pathogen infection and oxidative stress. J. Exp. Bot. 2008, 59, 2133–2146. [Google Scholar] [CrossRef] [PubMed]
  101. Asakura, Y.; Galarneau, E.; Watkins, K.P.; Barkan, A.; Van Wijk, K.J. Chloroplast RH3 DEAD box RNA helicases in maize and Arabidopsis function in splicing of specific group II introns and affect chloroplast ribosome biogenesis. Plant Physiol. 2012, 159, 961–974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Hok, S.; Danchin, E.G.; Allasia, V.; Panabieres, F.; Attard, A.; Keller, H. An Arabidopsis (malectin-like) leucine-rich repeat receptor-like kinase contributes to downy mildew disease. Plant Cell Environ. 2011, 34, 1944–1957. [Google Scholar] [CrossRef] [PubMed]
  103. Bruggmann, R.; Abderhalden, O.; Reymond, P.; Dudler, R. Analysis of epidermis-and mesophyll-specific transcript accumulation in powdery mildew-inoculated wheat leaves. Plant Mol. Biol. 2005, 58, 247–267. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Zheng, L.; Huang, J.; Yu, D. Isolation of genes expressed during compatible interactions between powdery mildew (Blumeria graminis) and wheat. Physiol. Mol. Plant Pathol. 2008, 73, 61–66. [Google Scholar] [CrossRef]
Figure 1. Frequency distribution of the F7 families derived from B73 (susceptible) × Ki11 (resistant) based on downy mildew (DM) incidence in Phnom Penh, Cambodia in September 2017. The x-axis indicates the frequency of DM incidence in F7 families. Ki11 and B73 were evaluated in April and September 2015.
Figure 1. Frequency distribution of the F7 families derived from B73 (susceptible) × Ki11 (resistant) based on downy mildew (DM) incidence in Phnom Penh, Cambodia in September 2017. The x-axis indicates the frequency of DM incidence in F7 families. Ki11 and B73 were evaluated in April and September 2015.
Genes 11 00191 g001
Figure 2. The linkage map of the F7 families of the cross between B73 (susceptible) and Ki11 (resistant). For each chromosome, the chromosome number is shown at the top, the markers on the right side, and the genetic distances in cM calculated using the Kosambi function on the left side. ★, a quantitative trait locus (QTL) position.
Figure 2. The linkage map of the F7 families of the cross between B73 (susceptible) and Ki11 (resistant). For each chromosome, the chromosome number is shown at the top, the markers on the right side, and the genetic distances in cM calculated using the Kosambi function on the left side. ★, a quantitative trait locus (QTL) position.
Genes 11 00191 g002
Figure 3. Relative expression levels (log2DM-infected/control) of 10 P. sorghi, P. maydis, and S. macrospora related genes in the control and infected plants measure via quantitative real-time polymerase chain reaction (qRT-PCR). The analysis was conducted on three independent plants as biological replicates. The data are presented as the mean ± standard error (SE). Student’s t-tests were used in the statistical analysis (* p < 0.05; ** p < 0.01; *** p < 0.001).
Figure 3. Relative expression levels (log2DM-infected/control) of 10 P. sorghi, P. maydis, and S. macrospora related genes in the control and infected plants measure via quantitative real-time polymerase chain reaction (qRT-PCR). The analysis was conducted on three independent plants as biological replicates. The data are presented as the mean ± standard error (SE). Student’s t-tests were used in the statistical analysis (* p < 0.05; ** p < 0.01; *** p < 0.001).
Genes 11 00191 g003
Figure 4. Relative expression levels (log2DM-infected/control) of 15 candidate genes for P. sorghi, P. maydis, and S. macrospora via qRT-PCR in the control and infected plants. The analysis was conducted on three independent plants as biological replicates. The data are presented as the mean ± SE values. Student’s t-tests were carried out in the statistical analysis (* p < 0.05; ** p < 0.01; *** p < 0.001).
Figure 4. Relative expression levels (log2DM-infected/control) of 15 candidate genes for P. sorghi, P. maydis, and S. macrospora via qRT-PCR in the control and infected plants. The analysis was conducted on three independent plants as biological replicates. The data are presented as the mean ± SE values. Student’s t-tests were carried out in the statistical analysis (* p < 0.05; ** p < 0.01; *** p < 0.001).
Genes 11 00191 g004
Table 1. Downy mildew (DM) Incidence of resistant (CML228, Ki3, and Ki11) and susceptible (B73, and CML270) parents in Phnom Penh, Cambodia 2015.
Table 1. Downy mildew (DM) Incidence of resistant (CML228, Ki3, and Ki11) and susceptible (B73, and CML270) parents in Phnom Penh, Cambodia 2015.
VarietyMid-April to JuneEarly September to October
4 Weeks6 Weeks4 Weeks6 Weeks
Resistant Genotype
CML2280%0%20%25%
Ki30%0%16.6%22.2%
Ki110%5%10%25%
Susceptible Genotype
B7350%100%100%100%
CML270100%100%75%100%
0%, highly resistant (HR); 1–10%, resistant (R); 11–25%, moderately resistant (MR); 26–50%, moderately susceptible (MS); 51–75%, susceptible (S); 76–100%, highly susceptible (HS).
Table 2. Statistical data for DM experiments conducted in Phnom Penh, Cambodia in September 2017.
Table 2. Statistical data for DM experiments conducted in Phnom Penh, Cambodia in September 2017.
TraitSample SizeMean of DM IncidenceStandard DeviationMinimumMaximumRangeSkewnessKurtosis
DM19286.4923.8701000–100−2.355.14
Table 3. Quantitative trait loci (QTLs) for P. sorghi, P. maydis, and S. macrospora resistance for the F7 families derived from B73 (susceptible) × Ki11 (resistant).
Table 3. Quantitative trait loci (QTLs) for P. sorghi, P. maydis, and S. macrospora resistance for the F7 families derived from B73 (susceptible) × Ki11 (resistant).
QTLsChr 1Bin LocationPositionLeft CIRight CILeft MarkerRight MarkerLOD 2PVE (%) 3Add 4Donor of DM-Resistant Genotype
qDM122.0111.0310.5311.53umc1165bnlg129714.1212.95−32.87Ki11
qDM22.0255.0354.5355.53umc2363phi0983.600.75−13.93Ki11
qDM333.0440.6040.1041.10umc1030phi2439663.210.48−8.39Ki11
qDM43.0580.6080.1081.10mmc0022bnlg42018.162.90−31.68Ki11
qDM566.05/6.0698.8098.3099.30umc1859bnlg34510.772.85−33.18Ki11
qDM699.0588.5888.0890.08umc1231umc23433.840.77−11.09Ki11
qDM79.07116.58115.08117.08dupssr29umc15053.170.47−8.27Ki11
1 Chr: chromosome; 2 LOD: logarithm of odds; 3 PVE: phenotype variance explained; 4 Add: additive effects by the alleles of Ki11 and B73.
Table 4. List of the DM-related genes in Arabidopsis thaliana, Oryza sativa, and Zea mays obtained from the NCBI Database (2019).
Table 4. List of the DM-related genes in Arabidopsis thaliana, Oryza sativa, and Zea mays obtained from the NCBI Database (2019).
Gene IDCultivarChromosomeLocationDescription
EDM2A. thaliana522,447,937–22,454,805Enhanced downy mildew 2
SGT1BA. thaliana46,851,184–6,853,912SGT1b
HSKA. thaliana27,508,473–7,509,887Homoserine kinase; downy mildew resistance 1
DMR6A. thaliana58,378,759–8,383,401Putative 2OG-Fe(II) oxygenase
LOC4324025O. sativa117,932,375–17,947,080Enhanced downy mildew 2
LOC107275878O. sativa38,811,797–8,818,014Enhanced downy mildew 2
LOC107275863O. sativa815,057,975–15,070,064Enhanced downy mildew 2-like
LOC4345309O. sativa815,114,902–15,130,493Enhanced downy mildew 2
LOC4345959O. sativa824,800,574–24,812,015Enhanced downy mildew 2
LOC4351808O. sativa126,760,146–6,763,038Enhanced downy mildew 2
LOC103647182Z. mays2160,527,731–160,580,975Enhanced downy mildew 2
LOC103648264Z. mays2241,693,442–241,703,192Enhanced downy mildew 2
LOC100191339Z. mays3135,359,547–135,361,124Downy mildew resistance 6
LOC103650325Z. mays3102,996,061–102,998,690Enhanced downy mildew 2
LOC100382073Z. mays4173,897,007–173,895,527Downy mildew resistance 6
LOC103654479Z. mays4202,734,383–202,764,701Enhanced downy mildew 2
IDC1Z. mays620,483,352–20,484,802Iron deficiency candidate 1; downy mildew resistance 6
LOC103632498Z. mays779,090,176–79,133,359Enhanced downy mildew 2
LOC103642860Z. mays10135,816,580–135,824,835Flavanone 3-dioxygenase 2
Table 5. List of the P. sorghi, P. maydis, and S. macrospora resistance candidate genes between the left and right flanked markers.
Table 5. List of the P. sorghi, P. maydis, and S. macrospora resistance candidate genes between the left and right flanked markers.
Transcript IDChr 1Bin LocationTypeLength (bp)Predicted Protein Size (aa)DescriptionProtein Information
AC210003.2_FG00422.01/2.02T01999332Uncharacterized LOC100274427Peroxidase 16
GRMZM2G020043T011938260--
GRMZM2G039345T01697206Ribulose bisphosphate carboxylase/oxygenase activase 2, chloroplasticP-loop containing nucleoside triphosphate hydrolase superfamily protein
GRMZM2G045049T012111521Probable phosphoribosylformylglycinamidine synthase, chloroplastic/mitochondrialProbable phosphoribosylformylglycinamidine synthase, chloroplastic/mitochondrial
GRMZM2G314171T011395464--
GRMZM2G363066T011450422G-type lectin S-receptor-like serine/threonine-protein kinase At1g34300G-type lectin S-receptor-like serine/threonine-protein kinase At1g34300
GRMZM2G13370733.04T01217---
GRMZM2G04767766.05/6.06T011199271Uncharacterized LOC100216590Abscisic acid receptor PYL5; abscisic acid receptor PYL3; pyrabactin resistance-like protein
GRMZM2G062031T011809382Uncharacterized LOC100276496Uncharacterized LOC100276496
GRMZM2G128315T013343964Probable LRR receptor-like serine/threonine-protein kinase IRKPutative leucine-rich repeat receptor-like protein kinase family protein
GRMZM2G00598499.05T01951222Photosystem II 11 kd proteinPhotosystem II repair protein PSB27-H1 chloroplastic; photosystem II protein
AC191071.3_FG0019.07T011605534Probable flavin-containing monooxygenase 1Probable flavin-containing monooxygenase 1
GRMZM2G028643T011954523Putative leucine-rich repeat receptor-like serine/threonine-protein kinase At2g14440Putative leucine-rich repeat receptor-like serine/threonine-protein kinase At2g14440; Leucine-rich repeat (LRR) family protein
GRMZM2G178880T011894574Uncharacterized LOC100191890Putative mannan synthase 7
GRMZM2G330907T013005759Uncharacterized LOC541659Leucine-rich repeat transmembrane protein kinase 3; leucine-rich transmembrane protein kinase2; putative STRUBBELIG family receptor protein kinase
1 Chr: chromosome.
Table 6. Pfam (EBI 2019) Domain Analysis of the 15 Upregulated Genes via qRT-PCR.
Table 6. Pfam (EBI 2019) Domain Analysis of the 15 Upregulated Genes via qRT-PCR.
Transcript IDFamilyDescription
AC210003.2_FG004PeroxidasePeroxidase (PF00141)
GRMZM2G020043HydrolaseHaloacid dehalogenase-like hydrolase (PF00702)
GRMZM2G039345--
GRMZM2G045049AIRS_C, GATase_5AIR synthase related protein, C-terminal domain (PF13507); CobB/CobQ-like glutamine amidotransferase domain (PF13507)
GRMZM2G314171DEAD, Metalloenzyme, MetalloenzymeDEAD/DEAH box helicase (PF00270); Metalloenzyme superfamily (PF01676); Metalloenzyme superfamily (PF01676)
GRMZM2G363066Pkinase_TyrProtein tyrosine kinase (PF07714)
GRMZM2G133707--
GRMZM2G047677Polyketide_cyc2Polyketide cyclase/dehydrase and lipid transport (PF10604)
GRMZM2G062031Microtub_bdMicrotubule binding (PF16796)
GRMZM2G128315LRRNT_2Leucine rich repeat N-terminal domain (PF08263)
GRMZM2G005984PSII_Pbs27Photosystem II Pbs27 (PF13326)
AC191071.3_FG001FMO-likeFlavin-binding monooxygenase-like (PF00743)
GRMZM2G028643Malectin_like, LRR_1Malectin-like domain (PF00560); Leucine rich repeat (PF00560)
GRMZM2G178880Glyco_trans_2_3Glycosyl transferase family group 2 (PF13632)
GRMZM2G330907LRRNT_2, LRR_8, LRR_1, LRR_8, PkinaseLeucine rich repeat N-terminal domain (PF08263); Leucine rich repeat (PF13855); Leucine rich repeat (PF00560); Leucine rich repeat (PF13855); Protein kinase domain (PF00069)

Share and Cite

MDPI and ACS Style

Kim, H.C.; Kim, K.-H.; Song, K.; Kim, J.Y.; Lee, B.-M. Identification and Validation of Candidate Genes Conferring Resistance to Downy Mildew in Maize (Zea mays L.). Genes 2020, 11, 191. https://doi.org/10.3390/genes11020191

AMA Style

Kim HC, Kim K-H, Song K, Kim JY, Lee B-M. Identification and Validation of Candidate Genes Conferring Resistance to Downy Mildew in Maize (Zea mays L.). Genes. 2020; 11(2):191. https://doi.org/10.3390/genes11020191

Chicago/Turabian Style

Kim, Hyo Chul, Kyung-Hee Kim, Kitae Song, Jae Yoon Kim, and Byung-Moo Lee. 2020. "Identification and Validation of Candidate Genes Conferring Resistance to Downy Mildew in Maize (Zea mays L.)" Genes 11, no. 2: 191. https://doi.org/10.3390/genes11020191

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

Article Metrics

Back to TopTop