Next Article in Journal
Exploring the Molecular Landscape of Nitrogen Use Efficiency in Potato (Solanum tuberosum L.) under Low Nitrogen Stress: A Transcriptomic and Metabolomic Approach
Previous Article in Journal
Effect of Chinese Milk Vetch on Zinc Content and Zinc Absorption of Rice in Purple Tidal Mud Soil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Assessing Heat Stress Tolerance of Wheat Genotypes through Integrated Molecular and Physio-Biochemical Analyses

1
Departmen Plant Production, College of Food and Agriculture Sciences, King Saud University, Riyadh 11451, Saudi Arabia
2
Department of Agricultural Engineering, College of Food and Agriculture Sciences, King Saud University, Riyadh 11451, Saudi Arabia
*
Author to whom correspondence should be addressed.
Agronomy 2024, 14(9), 1999; https://doi.org/10.3390/agronomy14091999
Submission received: 20 July 2024 / Revised: 24 August 2024 / Accepted: 27 August 2024 / Published: 2 September 2024
(This article belongs to the Section Crop Breeding and Genetics)

Abstract

:
Heat as an abiotic stress significantly impairs the sustainable productivity of wheat (Triticum aestivum L.). To determine the tolerance of genotypes to heat stress, a comprehensive approach should be used that integrates simultaneous phenotyping and genotyping analyses. The aim of this study is to identify local heat-tolerant genotypes using simple sequence repeat (SSR) markers and evaluate the selected genotypes under field conditions for their tolerance to heat stress. Of the 12 SSR markers that showed polymorphism, eight were associated with six important traits. The use of hierarchical cluster analysis (HC) based on SSR markers led to the identification of 13 genotypes that showed varying results and were grouped into three distinct heat tolerance classes: tolerant (T), moderately tolerant (MT), and sensitive (S). The results showed that heat stress had a significant effect on 19 traits under this study, with significant variation in tolerance to heat stress between genotypes. The tolerant genotypes exhibited a range of average thousand-kernel weight (TKW) values between 40.56 and 44.85, while the sensitive genotype (Yecora Rojo) had an average TKW of 35.45. Furthermore, the tolerant genotypes showed two to three times higher levels of antioxidants compared to the sensitive genotypes when exposed to heat stress. Among the traits analyzed, six showed a favorable combination of high heritability (>60%) and genetic gain (>20%). Through the integration of principal component analysis and stepwise multiple linear regression, it was determined that six traits (grain yield, 1000-kernel weight, plant height, intercellular carbon dioxide, flag leaf area, and grain filling duration) revealed differences between the 13 genotypes. HC analysis of the six traits resulted in the same division of genotypes into three main categories as observed in an HC analysis based on SSR markers. It is worth noting that Saudi wheat, including KSU106, KSU105, and KSU115 as local genotypes, in addition to the 16HTWYT-22 genotype, showed higher heat tolerance compared to the other genotypes tested, indicating its potential suitability for agriculture in Saudi Arabia. These results contribute to breeding programs focused on developing heat-tolerant wheat varieties and accelerate progress in wheat productivity improvement programs.

1. Introduction

Wheat (Triticum aestivum L.) is a cereal crop of great importance to world agriculture, but its agriculture often suffers from the harmful effects of heat stress. In Saudi Arabia, for instance, a rise in temperature by one degree Celsius can reduce crop productivity by 7–25% [1]. According to the 2014 Intergovernmental Panel on Climate Change (IPCC) report, there has been a consistent rise in atmospheric temperatures since the beginning of the 21st century, with projections of a further increase of approximately 1.0–1.7 °C [2]. Consequently, wheat growth and development have been increasingly challenged by more frequent and severe heat stress episodes as global climate change continues to unfold [3]. In connection with this, World Bank’s 2019 report indicated that the global population is growing at a rate of 1.1% per year [4]. As a result, the demand for food to feed this growing population is increasing. Thus, studies on wheat have attracted considerable attention in recent years. However, research regarding Saudi wheat genotypes, particularly their resistance to climate change, is very limited.
Climate factors, especially heat, play a pivotal role in shaping agricultural productivity and are recognized as one of the most critical abiotic stresses. When temperatures surpass a specific threshold, it leads to a phenomenon known as “heat stress” [5]. Heat stress occurring during the plant reproductive phase has crucial effects on both grain yield and quality by reducing the duration and activity of leaf photosynthesis, as well as altering the metabolism of sugars [6,7,8,9]. When exposed to severe heat stress, the metabolism of plants can incur detrimental effects due to the generation of reactive oxygen species (ROS) during metabolism, leading to the peroxidation of lipid membranes and the impairment of nucleic acids and proteins [10,11]. Heat stress shortens the period of canopy growth and reduces the plant’s leaf area, affecting the plant’s ability to intercept radiation and exchange carbon dioxide through the stomata. Stomatal closure and decreased activity of the Rubisco enzyme (which is crucial for photosynthesis) disrupts the photosynthetic process. As a result, the plant’s biomass accumulation and grain-filling period are reduced [12,13]. To alleviate these consequences, plants activate both enzymatic and non-enzymatic ROS-scavenging systems to neutralizing the damaging impact of ROS [14]. These systems encompass enzymes such as catalase (CAT), peroxidase (POD), and superoxide dismutase (SOD). Plants have been observed to exhibit enhanced tolerance to heat stress when their antioxidant capacity is increased [15,16]. Grasses that had acclimated to heat environment were found to have lower levels of ROS due to their higher antioxidant concentrations [17]. The increase in antioxidant capacity within plant cells may be triggered by certain signaling molecules [18,19,20]. This suggests that the synthesis, accumulation, and upregulation of antioxidant pathways are adaptive responses employed by plants to mitigate the damaging effects of heat stress [21,22,23]. This adaptive response helps protect plant cells from the damaging effects of ROS [24,25,26]. Studies on wheat genotypes have revealed a positive correlation between antioxidant enzyme levels and chlorophyll content, as well as a negative correlation between antioxidant enzymes and the membrane injury index across most growth stages [15,27]. This observation can be attributed to the ability of antioxidants to scavenge ROS, maintain chlorophyll and the cell membrane, and protect leaves from senescence, thereby ensuring the continuity of metabolism and maintenance of plant productivity [28,29]. Physiological traits associated with heat tolerance, such as antioxidant capacity, chlorophyll content, proline, glycine betaine, and leaf water content, can serve as valuable indirect selection tools for genetically improving wheat cultivars [30,31,32]. These physiological indicators tend to be less affected by environmental factors and exhibit greater genetic stability compared to traits directly associated with yield [33,34,35,36].
Heat tolerance is a multifaceted genetic trait governed by many genes, which makes the task of selecting and developing genotypes that exhibit both high productivity and heat tolerance complex. This complexity poses significant challenges for plant breeders in achieving sustainable food production [37,38]. Consequently, plant breeders face major challenges in achieving this goal, necessitating close collaboration with researchers and experts in related fields [39]. Molecular marker techniques have been widely used in genetic studies, such as molecular-assisted selection (MAS) and locus mapping quantitative traits (QTL) [40,41,42,43], with several genome-wide association studies (GWAS) focusing on agronomic traits and abiotic tolerance to establish links between phenotypic and genetic differences [44,45]. Laboratory tests provide more accurate data to evaluate heat tolerance in genotypes, due to the strong correlation between resistance to abiotic stresses observed in both field and laboratory conditions [46,47,48,49]. DNA markers offer the advantages of independence from environmental conditions, developmental stages, and organ types, thus enhancing the reliability of results [50,51,52,53]. Among various molecular markers, simple sequence repeat (SSR) markers are of great interest as they show high information, specificity, and neutrality, and they possess multiple allelic characteristics. They are widely distributed across the genome, show reproducibility in different environments, and possess common and high-throughput properties [54,55,56]. This would improve the breeding efficiency of yielding and heat-tolerant genotypes [43,57,58]. Therefore, it is essential to identify molecular markers associated with trait-specific genes [42,43,59,60,61,62,63,64,65,66]. Extensive research has been dedicated to identifying the genomic regions associated with heat tolerance in plants. These studies have utilized QTL analysis, which examines the relationship between genetic markers and observable traits indicative of heat tolerance. The key traits investigated include grain filling duration, thousand-grain weight, yield, and canopy temperature [67,68,69], as well as grain size, milling quality in wheat [70,71], and oil, starch, and protein concentrations in maize [72]. Additionally, researchers have studied traits related to greening and senescence [69,73], thylakoid membrane damage, plasma membrane damage, and chlorophyll content [33,74,75,76], as well as grain weight stability associated with the stay-green trait [66,77]. Acuna-Galindo et al. [65] identified 66 QTL associated with 81 different traits related to drought and heat stress adaptation. Numerous QTLs associated with heat tolerance have been identified across various developmental stages in plants [78,79].
Recent researches efforts have focused on the use of molecular markers and the examination of morphological, physiological, and biochemical parameters to select superior cultivars [80,81]. It is essential to have inexpensive, rapid, and easily measurable methods for detecting heat tolerance in a large number of genotypes [46,82]. Plant breeders and researchers often employ a range of multivariate analysis techniques and multidimensional methods, such as PCA, SMLR, PC analysis, and HC, to screen for and differentiate the sources of variation in their data [80,81,82,83]. These statistical tools are crucial for accurately verifying and selecting desirable traits in breeding programs. To achieve success in developing heat-tolerant wheat cultivars, it is essential to integrate agro-physiological traits into the breeding process [84,85,86]. By combining insights gained from these complementary approaches, including phenotypic evaluation and molecular marker analysis, this study aimed to identify the best-performing, heat-tolerant, high-yielding wheat genotypes suitable for cultivation in the Saudi environment with the help of molecular markers and physio-biochemical features, in addition to multiple analyses. This integrative approach can provide a more comprehensive understanding of the complex genetic and physiological mechanisms underlying heat adaptation in wheat. Ultimately, this knowledge could enable the development of improved, heat-tolerant, high-yielding wheat varieties that are promising for future breeding programs.

2. Materials and Methods

Sixty genotypes were used to investigate molecular markers associated with heat tolerance. Thirteen genotypes (four tolerance, four moderate tolerant, and five sensitive) were selected for field and physiological studies based on the molecular results (Table S1).

2.1. Genomic DNA Extraction and SSR Markers

Sixty wheat genotypes were germinated in growth chambers. Once the plants reached the fourth leaf stage, the leaves were collected and immediately frozen in liquid nitrogen, then ground using a pestle in a mortar. Genomic DNA was extracted using a Cetyltrimethylammonium ammonium bromide (CTAB) procedure, as described by Saghai Maroof et al. [87]. The quantity and quality of DNA were estimated using 0.8% agarose gels. The DNA was stored at −20 °C for further analysis. Additionally, the quality and concentration of DNA were estimated using a Nano Drop Spectrophotometer (ND-8000, Thermo-scientific Wilmington, DE, USA). The DNA samples were then diluted to a concentration of 20 ng/μL using ddH2O and stored at −20 °C for SSR fingerprinting.
Forty SSR markers associated with heat stress were utilized in this examination, as established by multiple researchers. PCR reactions were conducted using a 96 thermocycler and following the protocol described in reference [88]. The PCR reactions were performed in a 20 μL volume using the Promega Green master mix, which contains dNTPs, Taq polymerase, and a 10× PCR buffer. The reaction mixture included 1 mM MgCl2, 15 pmol of each primer, and 100 ng of genomic DNA as the template. Following PCR amplification, the resulting products were subjected to electrophoresis on 3% agarose gels. Ethidium bromide (EtBr) was used to stain the gels and visualize the DNA fragments, which were then placed in a Gel doc system. A 100 base pair ladder served as a standard for size comparison. The DNA fragments that separated in the agarose gels were visually examined and scored. Each allele fragment was scored as either 1 (presence) or 0 (absence) for each marker. Out of the 40 evaluated SSR markers, 12 exhibited polymorphism. These markers were found to be associated with heat stress tolerance in 12 genotypes. The study evaluated 13 genotypes (DHL2, 16HTWYT-22, KSU115, Yecora Rojo, 16HTWYT-38, KSU105, Klassic, KSU106, 16HTWYT-30, 16HTWYT-20, 16HTWYT-9, 16HTWYT-12, and Line-47) in the field. The genotypes underwent various measurements and assessments as part of the study.

2.2. Phenotyping, and Data Recording

2.2.1. Experimental Design

Thirteen genotypes were planted at the College of Food and Agricultural Sciences farm in Riyadh, Saudi Arabia during the 2021/22 growing season. The experimental fields were planted on two different dates: November 25, representing normal (C) conditions, and December 25, representing heat stress (H) conditions. The genotypes were planted in rows measuring 2.0 m in length with a spacing of 0.20 m between them. The experimental design used a split-plot arrangement with three replicates. Sowing dates assigned the main plots, while different genotypes were represented in the sub-plots. Cultural practices were carried out following the recommended guidelines.

2.2.2. Agro-Physio-Biochemical Traits

This study examined 17 agro-physio-biochemical traits. The agronomic traits were measured using the methodology provided by Al-Ashkar et al. [30]. These traits included duration of grain filling (GFD, in days), grain filling rate (GFR in g/day), plant height (PH in cm), flag leaf area (FLA in cm), spikelets per spike (NSS per spike), spike length (SL in cm), number of grains per spike (NGS), number of spike per square meter (NS per m2), grain yield (GY ton ha−1), and thousand-kernel weight (TKW g). Additionally, we evaluated the physiological trait of net photosynthetic rate (Pn) and intercellular carbon dioxide (Ci) measured using a portable photosynthesis system Li-640 (LI-COR, Inc., Lincoln, NE, USA), in addition to several biochemical traits, including polyphenol oxidase (PPO), superoxide dismutase (SOD), catalase (CAT), peroxidase (POD), and glycine betaine (GB), as shown below.
The levels of antioxidant enzymes were measured, including catalase (CAT), peroxidase (POD), superoxide dismutase (SOD), and polyphenol oxidase (PPO). Fresh leaf samples weighing 0.5 g were utilized. The leaves were crushed in liquid nitrogen and suspended in a buffer solution consisting of 50 mM potassium phosphate buffer (pH 7.8) and 1% (w/v) polyvinyl polypyrrolidone for extraction. The samples were then centrifuged at 14,000 rpm for 10 min at a temperature of 4 °C. The enzyme extract used to assess the activity of CAT, POD, PPO, and SOD in subsequent assays was obtained from the supernatant, which was prepared according to the instructions provided in [89,90,91].
The CAT activity was measured following the protocol described by Aebi [89]. The assay was performed in a 3 mL reaction mixture containing the following components: 1 mL of 0.1 M PB (pH 7.2), 1 mL of 75 mM H2O2 solution, and 1 mL of the enzyme extraction. The decrease in absorbance at 240 nm was recorded every 20 s for a total duration of 3 min. The SOD activity was measured according to Kono [92]. A reaction mixture was prepared that contained the following components: 1.3 mL of 50 mM sodium carbonate buffer, 500 μL of 96 μM NBT, 100 μL of 0.6% Triton X-100, and 100 μL of 20 mM hydroxylamine hydrochloride. This reaction mixture was incubated at room temperature for 2 min; then, 70 μL of enzyme extract was added. The absorbance of the mixture was immediately read at 560 nm, reading the absorbance every 20 s for 1–2 min. Reaction components for POD 3.0 mL PB (100 mM, pH 7.0), 50 μL guaiacol solution (20 mM), 30 μL H2O2 solution (12.3 mM) and 100 μL of enzyme extract were mixed well, and the absorbance was read at 436 nm. The absorbance was read up to 3 min [93]. The PPO activity was determined using the method outlined by Duckworth and Coleman [91]. The reaction was carried out in a mixture containing 0.03 mL of the enzyme solution and 1.74 mL of a 20 mM catechol solution (was prepared using a 50 mM PB, pH 6.8 at 25 °C). The absorbance of the reaction mixture was then measured at a wavelength of 420 nm.
The antioxidant enzyme activities were all measured based on protein concentration of the samples. The protein quantification was carried out following the protocol described by Bradford [94], utilizing bovine serum albumin (BSA) as the calibration standard.
The levels of Glycine betaine (GB) were determined by analyzing ground leaf samples with liquid nitrogen, following the methodology described by Sallam et al. [81]. Subsequently, 1 mg of the ground sample was transferred to a glass tube, and 1.5 mL of 2 N H2SO4 was added. The tube was then placed in a water bath set at 60 °C for 10 min to extract Glycine betaine. After centrifugation at 14,000 rpm for 10 min, the resulting supernatants were collected for further analysis. To determine the concentration of GB, 125 μL of the supernatant sample was mixed with 50 μL of a cold solution of potassium tri-iodide (KI-I2). The KI-I2 solution was prepared by dissolving 15.7 g of iodine and 20 g of potassium iodide in 100 mL of distilled water. The mixture was left at a temperature of 0–4 °C for 16 h and then centrifuged at 14,000 rpm for 15 min. The upper liquid phase was discarded. To dissolve any remaining small crystals in the tube chamber, 1.4 mL of 1,2-dichloroethane was added, and the solution was incubated for approximately 2–2.5 h. The resulting samples were analyzed using a spectrophotometer at a wavelength of 365 nm (U-2000, Hitachi Instruments, Tokyo, Japan) to determine the concentration of GB. A standard curve was prepared using betaine stock solution with concentrations of 1, 2, 4, 6, and 8 μL, which served as references to calculate the GB concentration in the samples.

2.2.3. Quality Traits

Two traits (protein content (PC) and gluten index (GI)) were estimated as explained in detail by the Committee [95].

2.3. Statistical Analysis

For agro-phenotypic analysis, the data of various parameters were analyzed by analysis of variance (ANOVA) and genetic parameters 19 traits were implemented using SAS v9.2 software (SAS Institute, Inc., Cary, NC, USA). The variance (mean squares) of data for 19 traits was used to compute variance components that are used to compute genetic parameters (genetic variance (σ2G), residual variance (σ2e), phenotypic variance (σ2Ph), heritability (h2 %), genotypic coefficient of variability (G.C.V %), phenotypic coefficient of variability (Ph.C.V. %), genetic advance (GA), and genetic gain (GG)), as described by Al-Ashkar et al. [96]. Principal component analysis (PCA) was carried out based on data provided by the correlation matrix to find out the variable contributing the most to the variance and the components loading the most on the variables. PCA identified important traits that were located in the first two components. In SMLR (stepwise multiple linear regression), PC (path coefficient), (hierarchical cluster), and LD (liner discriminant) analysis, 19 traits were used in SMLR to determine the key traits that contribute to enhancing and developing the variable of interest (GY), after which PC analysis was used to divide variation into direct and indirect effects. The effective indices (six out of nineteen traits) were used in the HC analysis to evaluate the phenotypic dissimilarity matrix between 13 genotypes, characterized into three groups For the Euclidean distance and ward’s method of agglomeration, LD analysis was employed to validate the genotype tolerance categories (the six indices used as quantitative variables) with the three categories (as quantitative variables). In the case of genotyping analysis, SSR bands were scored (present (1) or absent (0) to create a binary matrix). The genetic dissimilarity (matrix of pairwise) between genotypes was calculated used the coefficient of Jaccard dissimilarity. Agglomerative HC analysis was implemented using the unweighted pair group average method (UPGAM). Statistical analyses (PCA, SMLR, PC, HC, and DFA) were implemented through XLSTAT statistical package software (vers. 2019.1, Excel Add-ins soft SARL, New York, NY, USA).

3. Results

3.1. Genotypic Analysis Based on SSR Markers

Based on the outcomes of the stepwise marker analysis, a hierarchical clustering (HC) approach was employed to categorize the 13 genotypes into three main groups. The clustering was performed using the marker that exhibited correlations with the parameters under control, heat stress, and relative change, as depicted in Table 1. The first category, denoted as “T” for tolerance, encompassed four genotypes (KSU105, KSU106, 16HTWYW-22, and KSU115); the second category, labeled as “MT” for moderate tolerant, included four genotypes (16HTWYW-38, 16HTWYT-20, 16HTWYT-9, and DHL2). Lastly, the third category, named “S” for sensitive, covered five genotypes (16HTWYW-30, 16HTWYW-12, Line47, Yecora Rojo, and Klassic) (Figure 1). The clustering of the 13 genotypes based on phenotypic distance was further evaluated for its association with genetic distance using the Mantel test. The test revealed a significant positive correlation (r = 0.440, p < 0.012, and alpha = 0.05) between the phenotypic distance and genetic distance. These positive correlations highlight the importance of utilizing SSR markers as an effective tool for identifying tolerant genotypes at the early stages of a breeding program.

3.2. Exploring the Relationship between SSR Markers and Agro-Physio-Biochemical Traits

The study employed stepwise multiple linear regression (SMLR) analysis to identify the SSR markers that had the strongest association with the agrophysical and biochemical traits measured under control, temperature stress conditions and relative rate of change (RC). The analysis revealed that specific SSR markers were significantly associated with 6 out of the 19 agrophysical and biochemical traits that were examined (Table 1). The cumulative R2 value under control conditions ranged from 0.328 for PH to 0.790 for GY; under heat conditions, it ranged from 0.357 for PH to 0.994 for GFD, and the RC value ranged from 0.656 for Ci to 0.965 for GFD. Interestingly, wmc326 was significantly correlated with all six traits under heat stress and under relative change except for plant height. It also showed significant correlation with Ci under all conditions and with FLA under heat stress and under relative change. The results also showed that six molecular markers (wmc326, wmc65, wmc54, wmc74, gwm337, and wmc527) were correlated with GFD under heat stress, control, and relative change and had the highest cumulative R2 0.994. The plant height trait was also associated with five markers (Table 1).

3.3. Phenotypic Analysis for Assessing Terminal Heat Tolerance in Wheat Genotyp

Significant variations (p < 0.01) were observed between the optimal and heat-stressed treatments (I) for 16 measured traits, as indicated by ANOVA. Similarly, highly significant variations (p < 0.01) were observed for the genotypes (G) and the interaction between treatment and genotype (I × G) across the same nineteen traits (Table 2). The (h2) for the traits exhibited high levels for twelve traits (60.592% ≤ h2 ≤ 93.872%) and moderate levels for six traits (59.396% ≤ h2 ≤ 42.134%). The GG values were high for seven traits (28.545% ≤ GG ≤ 20.452%) and moderate for seven traits (14.984% ≤ GG ≤ 10.567%). The PCV and GCV The values exhibited both convergence and divergence across different traits. moreover, for all traits, the σ2G was smaller than the σ2Ph (Table 2).

3.4. Principal Component Analysis of Agro-Physio-Biochemical Traits under Control and Heat Stress

The first three principal component analyses (PCAs) showed that the eigenvalues were greater than 1 and together explained 87.018% of the total variance for the 19 traits studied. PCA1 and PCA2 showed 63.752% and 16.861% of the total variance, respectively. PCA1 (values of ≥0.337) was associated with 16 traits (PPO, CAT, SOD, Pn, Ci, FLA, GFD, GFR, PH, SL, NGS, NSS, NS, TKW, GY, and PC), while PCA2 was primarily related to two traits (GB and POD), and PCA3 was specifically linked to the GI trait (Table 3). The correlations between traits on PCA1 and PCA2, as illustrated in Figure 2, are indicative of the relationships between these traits. The PCA analysis accounted for a total of 19 traits. PCA1 displayed a positive correlation with thirteen traits and a negative correlation with six traits. On the other hand, PCA2 exhibited a positive correlation seventeen traits and a negative correlation with two traits. It is noteworthy that all genotypes subjected to heat stress demonstrated negative correlations with PCA. Furthermore, seven of these genotypes exhibited negative correlations with PCA2 as well.

3.5. SMLR and PC Analysis for the Preformance of Yield Traits

To prioritize yield as the primary objective, both stepwise multiple linear regression (SMLR) and path coefficient (PC) analysis were employed to identify genetically influenced traits. The SMLR analysis yielded findings that were noteworthy with regard to the contributions of various traits. Among the traits under consideration, TKW, PH, and Ci demonstrated significant contributions, with values of 0.868, 0.066, and 0.025 respectively (Table 4). Collectively, these three traits accounted for 0.958 of the total contribution (with a residual value of 0.205), as shown in Table 4. Given the strong correlation observed between TKW and yield, further investigation was conducted into traits associated with TKW. Specifically, FLA exhibited a substantial contribution value of 0.780. Given the strong correlation between FLA and TKW, an additional analysis was conducted to explore the traits linked to FLA and determine their direct and indirect effects on the same trait. In this analysis, two traits were identified as strongly associated with leaf area: GFD and Ci, contributing 0.699 and 0.146, respectively. Collectively, these two traits accounted for 0.845 of the total contribution, with a residual value of 0.394.
Based on the results of the SMLR, the effects of the three traits (TKW, PH, and Ci) that showed an association with GY were divided into direct and indirect effects through the path coefficient analysis. The TKW trait affected the GY for 0.244 (as a direct effect) out of 0.958, while the total direct and indirect impacts for the three traits were 0.429 and 0.529, respectively. When FLA was considered as the dependent variable, GFD trait affected leaf area for 0.261 (as a direct effect) out of a total of 0.845, while the total direct and indirect effects of the two traits (GFD and Ci) were 0.513 and 0.332, respectively (Table 4).

3.6. Hierarchical Clustering and Linear Discriminant Analysis

Based on the results of the SMLR and PC (which showed that five traits were associated with the GY and contributed to significant effects on the GY, whether directly or indirectly), we conducted a hierarchical clustering analysis of the 13 genotypes based on their performance using RC for these five traits (TKW, PH, Ci, FLA, and GFD) in addition to the GY. The results showed that the hierarchical distribution of the genotypes was distributed into three distinct groups. The first category, labeled as “T” for tolerant, encompassed four genotypes (KSU106, KSU105, 16HTWYT-22, and KSU115). The second category, denoted as “S” for sensitive, included seven genotypes (16HTWYT-30, Yecora Rojo, 16HTWYT-20, Klassic, 16HTWYT-12, Line47, and 16HTWYT-9). Finally, the third category, named “MT” for moderate tolerant, consisted of two genotypes (DHL2 and 16HTWYT-38) (Figure 3). To increase the credibility of the categorizations, an LD analysis was performed. This analysis aimed to validate the categories of the three groups (T, MT, and S) both before and after the analysis. The results of the LD analysis showed that the prior and posterior categories were completely consistent across all 13 genotypes, achieving a 100% accuracy rate. The membership probability values, which indicate the degree of compatibility between the prior and posterior categories, were (>0.5), with a value of 1 for all genotypes included in the study (Table 5).

4. Discussion

Abiotic stresses significantly impact plant growth and productivity, with heat and drought being the primary factors affecting global wheat production [97,98]. Climate projections indicate that global average temperatures are expected to rise by approximately 6 °C by the end of the 21st century [99,100]. Research has shown that even a 1 °C increase in temperature could result in a 6% decline in worldwide wheat production, particularly if the warming occurs during the plant’s reproductive phase, resulting in substantial losses in grain yield [101,102]. The integration of genotyping and molecular marker development holds great potential to enhance our understanding of these issues and provide valuable resources for selecting and breeding new wheat varieties that are tolerant to environmental stresses, including heat stress, which is particularly relevant for Saudi Arabia [40,103,104,105]. These advancements could play a critical role in ensuring sustainable wheat production in the face of a changing climate. The surrounding environment can influence the genotypes, and variations in soil conditions can negatively impact the field evaluation of these genotypes. However, the DNA within the cells remains unaffected by these external factors. This distinction is what sets selection based on molecular markers apart from other approaches [25,106,107,108].
The marker trait association (MTA) approach represents a promising avenue for plant breeding, offering a potential solution to the limitations inherent in linkage mapping [109,110]. In the present study, a set of 60 genotypes was screened using an SSR marker. From this screening, 13 selected genotypes displayed varying degrees of heat tolerance. These molecular markers, which are linked to quantitative trait loci (QTLs) in Marker-Assisted Selection (MAS), play a crucial role in determining key agro-physio-biochemical traits under heat stress [61,62,111,112,113,114]. In this particular study, we conducted HC analysis based on SSR markers, resulting in the categorization of 13 genotypes into three distinct categories (Figure 1). By employing a complementary approach that combined phenotyping analysis (based on phenotypic distance) and genotyping analysis (based on genetic distance), we revealed a significant positive correlation (r = 0.440 and p < 0.012) through the Mantel test. The strong correlation observed between the phenotypic and genotypic traits of the genotypes studied underscores the close association between these characteristics. This is a significant finding, as it emphasizes the effectiveness of SSR markers as a valuable tool for identifying tolerant genotypes early on in a breeding program. The results of this study align with numerous previous reports that have documented similar findings [61,62,81,114]. Interestingly, in the genotyping analysis, two genotypes (16HTWYT-20 and 16HTWYT-9) exhibited greater genetic distance compared to the categories determined by phenotyping analysis (Figure 1 and Figure 2). However, out of the 13 genotypes assessed, 11 (16HTWYT-30, 16HTWYT-12, KSU106, KSU105, Line-47, DHL2, 16HTWYT-22, KSU115, 16HTWYT-38, Yecora Rojo, and Klassic) demonstrated complete concordance between genotyping and phenotyping analyses. The congruence lends credibility and precision to both techniques employed in the evaluation and sorting of genotypes, facilitating the selection of the optimal genotypes capable of tolerating varied environmental conditions.
Given the importance of molecular markers and their independence from the surrounding environment, their association with crop, physiological, and biochemical traits is of great importance and reliability for plant breeders. Our results showed a strong association between some SSR markers (wmc326, wmc65, wmc54, wmc74, gwm337, and wmc527) and six important traits (GY, TKW, FLA, PH, GFD, and Ci) (Table 1). These traits can be effective indicators in screening and selecting genotypes for their tolerance to heat stress and for future studies in breeding and improvement programs [33,61,66,67,115,116].
Because productivity under field conditions is the main goal of breeding programs, field evaluation of the 13 selected genotypes was conducted under heat stress and control conditions by evaluating 19 agro-physio-biochemical traits (Table 2). The results revealed significant differences in performance between the genotypes under the control and heat stress conditions, indicating that these genotypes were highly susceptible to the effects of heat stress (Table 2) [23,25]. Moreover, the genotypes exhibited a high degree of genetic diversity, with varying performance across all the measured traits. The heat-tolerant genotypes (KSU106, KSU105, 16HTWYW-22, and KSU115) displayed the highest mean values for the assessed traits. For instance, the heat-tolerant genotype KSU115 recorded the highest value of TKW, GY, and FLA of 42.20 g, 6.77 ton ha−1, and 54.9 cm2, respectively, compared to the sensitive genotype Yecora Rojo, which had 35.4 g, 4.7 ton ha−1, and 25.21 cm2, respectively, under the same heat stress conditions [57,81,84]. Furthermore, the heat-tolerant genotypes exhibited the highest mean values for antioxidants and glycine betaine. The mean values for antioxidants in the heat-tolerant genotypes were twice as high as those recorded in the sensitive genotypes [23,24].The findings indicate that heat-tolerant plant genotypes possess superior physiological and biochemical mechanisms to mitigate the detrimental effects of heat stress [117,118,119]. This enables them to outperform heat-sensitive genotypes. The differential response of heat-tolerant genotypes to heat stress is attributed to their enhanced capacity to produce antioxidants (CAT, POD, and SOD) when exposed to heat stress [27,109]. These antioxidants protect cellular membranes and proteins by scavenging ROS [24,26,120]. As a result, they preserve chlorophyll levels and leaf health, allowing photosynthesis to continue unimpeded. This helps maintain high wheat yields under stressful heat conditions, in contrast to heat-sensitive genotypes that are negatively impacted by the stress and lack sufficient antioxidant levels to provide adequate protection. This genetic diversity can be leveraged to develop effective phenotypic selection strategies for breeding heat-tolerant cultivars, particularly for quantitative traits as reported in previous studies [121,122,123]. The strong interaction observed between the genotypes and the environmental conditions was a key factor for the traits evaluated in this study. This highlights the necessity for conducting extensive evaluations of genotypes across a diverse range of environments in order to thoroughly assess and characterize their heat-related characteristics [124]. Phenotypic variation among the genotypes played a notable role in the dominant genotype–environment interaction for all traits, as evidenced by the findings presented in Table 2. The level of heat tolerance improvement depends on the extent of genetic variation and heritability of the traits. (h2) provides valuable information about the proportion of genetic contribution relative to phenotypic variation, enabling predictions of validity and the reliability of phenotypic values [80,125,126,127]. Previous studies by Al-Ashkar et al. [80] and Burton [128] have demonstrated that combining h2, (GCV), and (GA) allows for credible assessments of the expected GG through phenotypic selection.
The h2 values exceeding 60.0% and the GG values exceeding 20.0% together indicate that the variance is due to genetic factors, making traits with high values of this parameter effective and reliable indicators for the selection of genotypes tolerant to different environmental conditions [127]. The GY trait is highly environmentally sensitive and has low h2 and is not reliable for direct selection of genotypes. Therefore, combining other traits can serve as powerful indicators for indirect selection of improved grain yield under heat stress conditions. This integrated, multi-faceted approach enables plant breeders to make more well-informed and balanced decisions when selecting desirable genotypes for their breeding programs [80].
In our study, we found that six traits (CAT, GB, FLA, SL, NGS, and GI) had high h2 (>60.0%) and high GG (>20.0%), and seven traits (PPO, SOD, Pn, Ci, GFR, GFD and PH) had intermediate h2 (>30.0%) and GG (>10.0%). And the remaining traits displayed one of the following scenarios: low h2 coupled with low GG, intermediate heritability but low genetic gain, or vice versa (Table 2). These results indicate that the first and second set of traits are reliable screening indicators for selecting genotypes for tolerance to different environmental stresses because they have high to moderate h2 and high to moderate GG [127,129]. These traits demonstrate effectiveness and promise as screening criteria in breeding programs. Utilizing them can facilitate the selection of genotypes with stress tolerance, which can then be used to transfer these advantageous traits to other genotypes through breeding. Notably, all the measured traits demonstrated a genotypic coefficient of variation (GCV) lower than the phenotypic coefficient of variation (PCV), despite the preference of breeders for higher GCV values over PCV values [130].
The PCA is a valuable tool for identifying significant traits positioned within the first two components, as demonstrated by loadings ≥ 0.337 in Table 3. Genotypes that exhibited heat tolerance (KSU106, 16 HWY-22, KSU105, and KSU110) demonstrated notable correlations with the majority of physiological traits, as depicted in Figure 2. By integrating genetic parameters and PCA outcomes, we discovered five traits (CAT, GB, FLA, SL, and NGS) that exhibited high heritability and genetic gain values while being located within the first two components. Consequently, these traits serve as efficient screening criteria [80,82].
Heat tolerance is a complex genetic trait that is significantly influenced by environmental factors. Consequently, the sole reliance on GY as an indicator of genotypes’ heat tolerance index is inadequate [80]. To ensure the accuracy of our findings, we employed additional statistical analyses, specifically SMLR and PC analyses. These analyses provide valuable insights into the dependent correlations with independent variables [129,131]. The results of the SMLR and PC analyses demonstrated a strong association positive between TKW, PH, and Ci with GY, with an R2 value of 0.958 and a significance level of p < 0.047. Among these traits, TKW contributed the most (0.868), followed by PH (0.066) and Ci (0.025), as indicated in Table 4. Similar results were found in [132,133,134]. Given the substantial correlation between TKW and yield, we proceeded to investigate the traits associated with TKW (R2 = 0.780, p < 0.0001). Notably, FLA emerged as a trait strongly linked to TKW, with a contribution value of 0.780. Given the pivotal role of leaf area in determining the weight of a thousand grains, we included it as a dependent attribute in order to investigate the associated traits; this is consistent with [81,84,135,136]. The SMLR analysis demonstrated that GFD and Ci were significantly correlated with FLA (R2 = 0.845, p < 0.012). The contribution rates of GFD and Ci were 0.699 and 0.146, respectively. To distinguish between direct and indirect impacts, we proceeded with a PC analysis. The analysis yielded noteworthy results when considering grain yield. The TKW exhibited a direct effect of 0.244, indicating a significant correlation between TKW and its direct influence on GY. Moreover, when the dependent variable was the FLA trait, the GFD trait displayed a direct effect of 0.261. These direct effects provide valuable insights into the association between the respective traits [30,137].
Research has shown that identifying key traits influencing crop yield and selecting for those strongly correlated with grain yield can be an effective breeding strategy. However, simply estimating the association between yield and its components may not be enough to fully understand the importance of each component in determining final grain yield. More advanced statistical techniques, such as SMLR and PC analysis, can provide deeper insights into the relationships between independent variables and GY. These methods can decompose the associations into direct and indirect effects on GY [138,139]. Plant breeders have also employed path analysis to help identify the most useful traits to target as selection criteria for improving overall crop yield [140,141]. By utilizing these more sophisticated analytical approaches, breeders can gain a more nuanced understanding of the complex factors driving grain yield.
Our result, as shown in Table 4, reveals that SMLR and PC analysis identified five traits (TKW, FLA, GFD, Ci, and PH) that were positively associated with GY. These traits hold significant value, as they provide breeders with important insights into the overall productivity of a given genotype. The strong positive correlation we observed between thousand-kernel weight and yield aligns with findings from previous studies [142,143]. Multiple analyses (SMLR and PC) in our study also identified a strong positive association between FLA, TKW, and GFD, which identified the importance of FLA and its direct and indirect effect on GY (Table 4). The FL is a critically important tissue for plant growth and development. As the uppermost leaf containing high concentrations of chlorophyll, the flag leaf plays a vital role in light interception, photosynthesis, and the storage of assimilates [144,145]. These key physiological functions make the flag leaf essential for crop productivity and the accumulation of carbohydrates in the grains [146,147]. Research has shown that the flag leaf’s photosynthetic rate is positively correlated with grain yield determination in rice [148,149]. The leaf area of a plant significantly impacts its overall growth, development, productivity, and quality by influencing light capture [150,151]. In wheat, the length and width of the flag leaf have been found to be positively correlated with important yield components like the number of spikelets per plant and grain weight [152,153]. Similarly, in rice, positive correlations have been reported between leaf thickness, grain weight per inflorescence, and the number of spikelets per inflorescence [154]. Interestingly, we found that the traits identified by SMLR and PC analysis as positively associated with yield (Table 4) were the same traits for which molecular marker showed strong associations (Table 1), making them reliable indicators for selection for heat tolerance and a topic that opens up prospects for future breeding programs to tolerate different environmental conditions and improve quality. By comprehending and quantifying the contributions of these traits to grain yield, we can pave the way towards enhancing the yield potential of wheat under heat stress conditions [155,156].
To further explore the relationships among the 13 genotypes under study, we conducted a HC analysis incorporating the five traits (TKW, PH, Ci, FLA, and GFD) identified through the previous SMLR and PC analyses, in addition to GY. We used the relative rate of change for these traits, and the results showed that the genotypes were distributed into three groups based on their performance under heat stress conditions. The S category comprised seven genotypes, while the T and MT categories included four and two genotypes, respectively. It is noteworthy that the HC analysis has been frequently utilized by researchers for the purpose of ranking heat-tolerant wheat genotypes [122,128,129,157,158]. We performed LD analysis to validate the accuracy of the classes. LD analysis showed complete consistency between the preceding and subsequent classes among the 13 genotypes used (Table 5). Furthermore, cross-validation showed that the preceding and subsequent classes were identical. Thus, LD analysis represents a unique statistical tool, providing selection criteria for the accuracy and reliability of genotype resources with respect to heat tolerance [111,112,113,159,160,161,162].

5. Conclusions

This study used a multivariate analytical approach to investigate genotypic and phenotypic characteristics associated with heat tolerance in wheat genotypes. SSR markers, linked to quantitative trait loci (QTLs) in marker-assisted selection (MAS), play a crucial role in determining key agronomic physical and biochemical traits under thermal stress. The results identified six traits that exhibited a combination of high heritability (>60%) and genetic gain (>20%), suggesting significant potential for selection and improvement. The SMLR and PC analyses collectively identified six major traits (GY, TKW, FLA, GFD, PH, and Ci) that were strongly associated with molecular markers and effectively showed phenotypic differences between genotypes, which were then used as the basis for HC analysis. The HC analysis divided the 13 tested genotypes into three distinct classes, a classification fully supported by LD and SSR marker analysis. The Mantel test revealed significant associations between a various of morphological, physiological, and biochemical traits and the SSR genetic markers evaluated in this study. These results highlight the potential utility of these trait–marker associations as effective screening criteria for identifying heat-tolerant wheat genotypes. Based on the study’s findings, the researchers identify four wheat genotypes that exhibited exceptional tolerance to heat stress. These included three local genotypes (KSU105, KSU106, KSU115) and one global genotype (16HTWYT-22). Notably, these heat-tolerant genotypes displayed the highest values across all the measured traits, even under the heat stress conditions. For example, the KSU115 genotype recorded the greatest TKW of 42.20 g, the highest GY of 6.77 ton ha−1 and the largest FLA of 54.9 cm2. Additionally, these genotypes possessed a remarkable capacity to scavenge oxygen free radicals, attributed to their high antioxidant content, which was particularly evident under heat stress conditions. These genotypes are promising candidates for incorporated into future breeding programs aimed at further evaluation their tolerance to other abiotic stresses, such as drought and salinity, by examining relevant physiological parameters and studying the gene expression of genes associated with tolerance to environmental stresses.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/agronomy14091999/s1, Table S1. Names, pedigree and Source of the 60 bread wheat genotypes.

Author Contributions

Conceived and designed the experiments: M.S. and I.A.-A.; performed the experiments: M.S.; analyzed the data: I.A.-A., M.S. and A.A.-D.; morpho-physiological measurements: M.S., A.M.Z. and K.A.A.-G.; edited the manuscript: I.A.-A., A.G. and M.S.; final approval of the version to be published: I.A.-A., A.G. and A.A.-D. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Researchers Supporting Project Number (RSPD2024-R954), King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

All data are available in the article or Supplementary Materials.

Acknowledgments

The authors extend their appreciation to the Researchers Supporting Project Number (RSPD2024-R954), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Haque, M.I.; Khan, M.R. Impact of climate change on food security in Saudi Arabia: A roadmap to agriculture-water sustainability. J. Agribus. Dev. Emerg. Econ. 2022, 12, 1–18. [Google Scholar] [CrossRef]
  2. IPCC. Climate Change: Synthesis Report. In Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; IPCC: Geneva, Switzerland, 2014; 151p. [Google Scholar]
  3. Liu, B.; Asseng, S.; Wang, A.; Wang, S.; Tang, L.; Cao, W.; Zhu, Y.; Liu, L. Modelling the effects of post-heading heat stress on biomass growth of winter wheat. Agricultural 2017, 247, 476–490. [Google Scholar] [CrossRef]
  4. Oberle, B.; Bringezu, S.; Hatfield-Dodds, S.; Hellweg, S.; Schandl, H.; Clement, J. Global Resources Outlook: 2019; International Resource Panel, United Nations Envio: Paris, France, 2019. [Google Scholar]
  5. Akter, N.; Rafiqul Islam, M. Heat stress effects and management in wheat. A review. Agron. Sustain. Dev. 2017, 37, 1–17. [Google Scholar] [CrossRef]
  6. Zhang, C.X.; Feng, B.H.; Chen, T.T.; Fu, W.M.; Li, H.B.; Li, G.Y.; Jin, Q.Y.; Tao, L.X.; Fu, G.F. Heat stress-reduced kernel weight in rice at anthesis is associated with impaired source-sink relationship and sugars allocation. Environ. Exp. Bot. 2018, 155, 718–733. [Google Scholar] [CrossRef]
  7. Saini, H.; Aspinall, D. Abnormal sporogenesis in wheat (Triticum aestivum L.) induced by short periods of high temperature. Ann. Bot. 1982, 49, 835–846. [Google Scholar] [CrossRef]
  8. Giorno, F.; Wolters-Arts, M.; Mariani, C.; Rieu, I. Ensuring reproduction at high temperatures: The heat stress response during anther and pollen development. Plants 2013, 2, 489–506. [Google Scholar] [CrossRef]
  9. Ortiz, R.; Sayre, K.D.; Govaerts, B.; Gupta, R.; Subbarao, G.; Ban, T.; Reynolds, M. Climate change: Can wheat beat the heat? Agric. Ecosyst. Environ. 2008, 126, 46–58. [Google Scholar] [CrossRef]
  10. Bita, C.E.; Gerats, T. Plant tolerance to high temperature in a changing environment: Scientific fundamentals and production of heat stress-tolerant crops. Front. Plant Sci. 2013, 4, 273. [Google Scholar] [CrossRef]
  11. Chakraborty, U.; Pradhan, D. High temperature-induced oxidative stress in, role of antioxidants and amelioration of stress by chemical pre-treatments. J. Plant Interact. 2011, 6, 43–52. [Google Scholar] [CrossRef]
  12. Ashraf, M. Thermotolerance in plants: Potential physio-biochemical and molecular markers for crop improvement. Environ Exp. Bot. 2021, 186, 104454. [Google Scholar] [CrossRef]
  13. Edreira, J.I.R.; Otegui, M.E. Heat stress in temperate and tropical maize hybrids: Differences in crop growth, biomass partitioning and reserves use. Field Crop. Res. 2012, 130, 87–98. [Google Scholar] [CrossRef]
  14. Suzuki, N.; Koussevitzky, S.; Mittler, R.; Miller, G. ROS and redox signalling in the response of plants to abiotic stress. Plant Cell Environ. 2012, 35, 259–270. [Google Scholar] [CrossRef]
  15. Almeselmani, M.; Deshmukh, P.S.; Sairam, R.K.; Kushwaha, S.R.; Singh, T.P. Protective role of antioxidant enzymes under high temperature stress. Plant Sci. 2006, 171, 382–388. [Google Scholar] [CrossRef]
  16. Babu, R.N.; Devaraj, V.R. High temperature and salt stress response in French bean (Phaseolus vulgaris). Aust. J. Crop Sci. 2008, 2, 40–48. [Google Scholar]
  17. Xu, S.; Li, J.L.; Zhang, X.Q.; Wei, H.; Cui, L.J. Effects of heat acclimation pretreatment on changes of membrane lipid peroxidation, antioxidant metabolites, and ultrastructure of chloroplasts in two cool-season turfgrass species under heat stress. Environ. Exp. Bot. 2006, 56, 274–285. [Google Scholar] [CrossRef]
  18. Gong, M.; Chen, S.N.; Song, Y.Q.; Li, Z.G. Effect of calcium and calmodulin on intrinsic heat tolerance in relation to antioxidant systems in maize seedlings. Aust. J. Plant Physiol. 1997, 24, 371–379. [Google Scholar] [CrossRef]
  19. Dat, J.F.; Foyer, C.H.; Scott, I.M. Changes in salicylic acid and antioxidants during induced thermotolerance in mustard seedlings. Plant Physiol. 1998, 118, 1455–1461. [Google Scholar] [CrossRef]
  20. Apel, K.; Hirt, H. Reactive oxygen species: Metabolism, oxidative stress, and signal transduction. Annu. Rev. Plant Biol. 2004, 55, 373–399. [Google Scholar] [CrossRef] [PubMed]
  21. Almeselmani, M.; Deshmukh, P.; Sairam, R. High temperature stress tolerance in wheat genotypes: Role of antioxidant defence enzymes. Acta Agron. Hung. 2009, 57, 1–14. [Google Scholar] [CrossRef]
  22. Sarafraz-Ardakani, M.R.; Khavari-Nejad, R.A.; Moradi, F.; Najafi, F. Abscisic acid and cytokinin-induced osmotic and antioxidant regulation in two drought-tolerant and drought-sensitive cultivars of wheat during grain filling under water deficit in field conditions. Not. Sci. Biol. 2014, 6, 354–362. [Google Scholar] [CrossRef]
  23. Gupta, N.K.; Agarwal, S.; Agarwal, V.P.; Nathawat, N.S.; Gupta, S.; Singh, G. Effect of short-term heat stress on growth, physiology and antioxidative defence system in wheat seedlings. Acta Physiol. Plant. 2013, 35, 1837–1842. [Google Scholar] [CrossRef]
  24. Anjali Anand, A.A.; Shantha Nagarajan, S.N.; Pathak, P. Effect of high temperature on hydrogen peroxide scavenging enzymes during reproductive phase in aromatic rice cultivars. Indian J. Plant Physiol. 2006, 11, 427–431. [Google Scholar]
  25. Sharma, P.; Mehta, G.; Shefali Muthusamy, S.K.; Singh, S.K.; Singh, G.P. Development and validation of heat-responsive candidate gene and miRNA gene based SSR markers to analysis genetic diversity in wheat for heat tolerance breeding. Mol. Biol. Rep. 2021, 48, 381–393. [Google Scholar] [CrossRef] [PubMed]
  26. Das, K.; Roychoudhury, A. Reactive oxygen species (ROS) and response of antioxidants as ROS-scavengers during environmental stress in plants. Front. Environ. Sci. 2014, 2, 53. [Google Scholar] [CrossRef]
  27. Senn, M.E.; Grozeff, G.E.G.; Alegre, M.L.; Barrile, F.; De Tullio, M.C.; Bartoli, C.G. Effect of mitochondrial ascorbic acid synthesis on photosynthesis. Plant Physiol. Bioch 2016, 104, 29–35. [Google Scholar] [CrossRef] [PubMed]
  28. Chauhan, S. Physiological and Molecular Basis of Heat Tolerance with Emphasis on Oxidative Stress Metabolism in Wheat. Ph.D. Thesis, HNB Garhwal University, Srinagar, India, 2005. [Google Scholar]
  29. Balla, K.; Bencze, S.; Janda, T.; Veisz, O. Analysis of heat stress tolerance in winter wheat. Acta Agron. Hung. 2009, 57, 437–444. [Google Scholar] [CrossRef]
  30. Al-Ashkar, I.; Alotaibi, M.; Refay, Y.; Ghazy, A.; Zakri, A.; Al-Doss, A.J.A. Selection criteria for high-yielding and early-flowering bread wheat hybrids under heat stress. PLoS ONE 2020, 15, e0236351. [Google Scholar] [CrossRef] [PubMed]
  31. Pinto, R.S.; Lopes, M.S.; Collins, N.C.; Reynolds, M.P. Modelling and genetic dissection of staygreen under heat stress. Theor. Appl. Genet. 2016, 129, 2055–2074. [Google Scholar] [CrossRef]
  32. Pinto, R.S.; Reynolds, M.P.; Mathews, K.L.; McIntyre, C.L.; Olivares-Villegas, J.J.; Chapman, S.C. Heat and drought adaptive QTL in a wheat population designed to minimize confounding agronomic effects. Theor. Appl. Genet. 2010, 121, 1001–1021. [Google Scholar] [CrossRef]
  33. Barakat, M.N.; Saleh, M.; Al-Doss, A.A.; Moustafa, K.A.; Elshafei, A.A.; Al-Qurainy, F.H. Identification of New Ssr Markers Linked to Leaf Chlorophyll Content, Flag Leaf Senescence and Cell Membrane Stability Traits in Wheat under Water Stressed Condition. Acta Biol. Hung. 2015, 66, 93–102. [Google Scholar] [CrossRef]
  34. Liu, K.Y.; Xu, H.; Liu, G.; Guan, P.F.; Zhou, X.Y.; Peng, H.R.; Yao, Y.; Ni, Z.; Sun, Q.; Du, J. QTL mapping of flag leaf-related traits in wheat. Theor. Appl. Genet. 2018, 131, 839–849. [Google Scholar] [CrossRef]
  35. Rehman, H.U.; Tariq, A.; Ashra, F.I.; Ahmed, M.; Muscolo, A.; Basra, S.M.A.; Reynolds, M. Evaluation of Physiological and Morphological Traits for Improving Spring Wheat Adaptation to Terminal Heat Stress. Plants 2021, 10, 455. [Google Scholar] [CrossRef]
  36. Mondal, S.; Singh, R.P.; Crossa, J.; Huerta-Espino, J.; Sharma, I.; Chatrath, R.; Joshi, A.K. Earliness in wheat: A key to adaptation under terminal and continual high temperature stress in South Asia. Field Crop. Res. 2013, 151, 19–26. [Google Scholar] [CrossRef]
  37. Lamba, K.; Kumar, M.; Singh, V.; Chaudhary, L.; Sharma, R.; Yashveer, S.; Dalal, M.S. Heat stress tolerance indices for identification of the heat tolerant wheat genotypes. Sci. Rep. 2023, 13, 10842. [Google Scholar] [CrossRef] [PubMed]
  38. Hussain, S.; Jamil, M.; Napar, A.A.; Rahman, R.; Bano, A.; Afzal, F.; Gul Kazi, A.; Mujeeb Kazi, A. Heat stress in wheat and interdisciplinary approaches for yield maximization. In Plant-Environment Interaction: Responses Approaches to Mitigate Stress; John Wiley & Sons: Hoboken, NJ, USA, 2016; pp. 161–183. [Google Scholar]
  39. Al-Ashkar, I.; Romdhane, W.B.; El-Said, R.A.; Ghazy, A.; Attia, K.; Al-Doss, A. Agro-Physiologic Responses and Stress-Related Gene Expression of Four Doubled Haploid Wheat Lines under Salinity Stress Conditions. Biology 2021, 10, 56. [Google Scholar] [CrossRef]
  40. Kushanov, F.N.; Turaev, O.S.; Ernazarova, D.K.; Gapparov, B.M.; Oripova, B.B.; Kudratova, M.K.; Rafieva, F.; Khalilov, K.; Erjigitov, D.; Khidivov, M.; et al. Genetic Diversity, QTL Mapping, and Marker-Assisted Selection Technology in Cotton (Gossypium spp.). Front. Plant Sci. 2021, 12, 779386. [Google Scholar] [CrossRef]
  41. Cao, L.-Y.; Zhao, J.-G.; Zhan, X.-D.; Li, D.-L.; He, L.-B.; Cheng, S.-H. Mapping QTLs for heat tolerance and correlation between heat tolerance and photosynthetic rate in rice. Chin. J. Rice Sci. 2003, 17, 223. [Google Scholar]
  42. Al-Doss, A.; Barakat, M.; Moustafa, K.; Motawei, M.; Alamri, M.; Elshafei, A.; Sallam, M.S.; Al-Ashkar, I.M. QTL analysis for grain failing rate and related traits in doubled haploid population of wheat under heat stress conditions. In Proceedings of the Conference Plant and Animal Genomics, San Diego, CA, USA, 10–15 January 2016; Volume 24. [Google Scholar]
  43. Barakat, M.N.; Al-Doss, A.A.; Elshafei, A.A.; Moustafa, K.A. Bulked segregant analysis to detect quantitative trait loci (QTL) related to heat tolerance at grain filling rate in wheat using simple sequence repeat (SSR) markers. Afr. J. Biotechnol. 2012, 11, 12436–12442. [Google Scholar]
  44. Ward, B.P.; Brown-Guedira, G.; Kolb, F.L.; Van Sanford, D.A.; Tyagi, P.; Sneller, C.H.; Griffey, C.A. Genome-wide association studies for yield-related traits in soft red winter wheat grown in Virginia. PLoS ONE 2019, 14, e0208217. [Google Scholar] [CrossRef] [PubMed]
  45. Jamil, M.; Ali, A.; Gul, A.; Ghafoor, A.; Napar, A.A.; Ibrahim, A.M.H.; Naveed, N.H.; Yasin, N.A.; Kazi, A.M. Genome-wide association studies of seven agronomic traits under two sowing conditions in bread wheat. BMC Plant Biol. 2019, 19, 149. [Google Scholar] [CrossRef] [PubMed]
  46. Evans, R.; Skaggs, R.; Sneed, R. Stress day index models to predict corn and soybean relative yield under high water table conditions. Trans. ASAE 1991, 34, 1997–2005. [Google Scholar] [CrossRef]
  47. Kpoghomou, B.; Sapra, V.; Beyl, C.; Science, C. Screening for drought tolerance: Soybean germination and its relationship to seedling responses. J. Agron. 1990, 164, 153–159. [Google Scholar] [CrossRef]
  48. Urrea-Gómez, R.; Ceballos, H.; Pandey, S.; BahíaFilho, A.F.; León, L.A. A greenhouse screening technique for acid soil tolerance in maize. Am. Soc. Agron. 1996, 88, 806–812. [Google Scholar] [CrossRef]
  49. Richard, C.; Munyinda, K.; Kinkese, T.; Osiru, D.S. Genotypic variation in seedling tolerance to aluminum toxicity in historical maize inbred lines of Zambia. Agronomy 2015, 5, 200–219. [Google Scholar] [CrossRef]
  50. Ainsworth, C.; Beynon, J.; Buchanan-Wollaston, V. Techniques in Plant Biology: The Practical Manual; Wye College, University of London: London, UK, 1996. [Google Scholar]
  51. Al-Faifi, S.A.; Migdadi, H.M.; Algamdi, S.S.; Khan, M.A.; Ammar, M.H.; Al-Obeed, R.S.; Al-Thamra, M.; EL-Harty, E. Development, characterization and use of genomic SSR markers for assessment of genetic diversity in some Saudi date palm (Phoenix dactylifera L.) cultivars. Electron. J. Biotechnol. 2016, 21, 18–25. [Google Scholar] [CrossRef]
  52. Khierallah, H.; Bader, S.; Baum, M.; Hamwieh, A. Assessment of genetic diversity for some Iraqi date palms (Phoenix dactylifera L.) using amplified fragment length polymorphisms (AFLP) markers. Afr. J. Biotechnol. 2011, 10, 9570–9576. [Google Scholar] [CrossRef]
  53. Hasan, N.; Choudhary, S.; Naaz, N.; Sharma, N.; Laskar, R.A. Recent advancements in molecular marker-assisted selection and applications in plant breeding programmes. J. Genet. Eng. Biotechnol. 2021, 19, 128. [Google Scholar] [CrossRef]
  54. Powell, W.; Machray, G.C.; Provan, J. Polymorphism revealed by simple sequence repeats. Trends Plant Sci. 1996, 1, 215–222. [Google Scholar] [CrossRef]
  55. Chen, X.; Min, D.; Yasir, T.A.; Hu, Y.G. Genetic diversity, population structure and linkage disequilibrium in elite Chinese winter wheat investigated with SSR markers. PLoS ONE 2012, 7, e44510. [Google Scholar] [CrossRef]
  56. Singh, A.K.; Chaurasia, S.; Kumar, S.; Singh, R.; Kumari, J.; Yadav, M.C.; Singh, N.; Gaba, S.; Jacob, S.R. Identification, analysis and development of salt responsive candidate gene based SSR markers in wheat. BMC Plant Biol. 2018, 18, 249. [Google Scholar] [CrossRef] [PubMed]
  57. Tadesse, W.; Ogbonnaya, F.C.; Jighly, A.; Sanchez-Garcia, M.; Sohail, Q.; Rajaram, S.; Baum, M. Genome-Wide Association Mapping of Yield and Grain Quality Traits in Winter Wheat Genotypes. PLoS ONE 2015, 10, e0141339. [Google Scholar] [CrossRef]
  58. Semenov, M.A.; Halford, N.G. Identifying target traits and molecular mechanisms for wheat breeding under a changing climate. J. Exp. Bot. 2009, 60, 2791–2804. [Google Scholar] [CrossRef]
  59. Beecher, F.W.; Mason, E.; Monda, L.S.; Awika, J.; Hays, D.; Ibrahim, A. Identification of quantitative trait loci (QTLs) associated with maintenance of wheat (Triticum aestivum Desf.) quality characteristics under heat stress conditions. Euphytica 2012, 188, 361–368. [Google Scholar] [CrossRef]
  60. Najeb Barakat, M.; Al-Doss, A.; Elshafei, A.A.; Moustafa, K.A. Identification of new microsatellite marker linked to the grain filling rate as indicator for heat tolerance genes in F2 wheat population. Aust. J. Crop Sci. 2011, 5, 104–110. [Google Scholar]
  61. Li, M.; Feng, J.; Zhou, H.; Najeeb, U.; Li, J.; Song, Y.; Zhu, Y. Overcoming Reproductive Compromise Under Heat Stress in Wheat: Physiological and Genetic Regulation, and Breeding Strategy. Front. Plant Sci. 2022, 13, 881813. [Google Scholar] [CrossRef]
  62. Sun, L.; Wen, J.; Peng, H.; Yao, Y.; Hu, Z.; Ni, Z.; Sun, Q.; Xin, M. The genetic and molecular basis for improving heat stress tolerance in wheat. Abiotech 2022, 3, 25–39. [Google Scholar] [CrossRef]
  63. Jha, U.C.; Bohra, A.; Singh, N.P. Heat stress in crop plants: Its nature, impacts and integrated breeding strategies to improve heat tolerance. Plant Breed. 2014, 133, 679–701. [Google Scholar] [CrossRef]
  64. More, K.; Achyut, B.; Jadhav, B.; Malge, H. Screening of Wheat (Triticum aestivum L.) genotypes for Heat Tolerance Trait by physio-molecular Markers. Biol. Forum Int. J. 2022, 14, 63–68. [Google Scholar]
  65. Acuña-Galindo, M.A.; Mason, R.E.; Subramanian, N.K.; Hays, D.B. Meta-analysis of wheat QTL regions associated with adaptation to drought and heat stress. Crop Sci. 2015, 55, 477–492. [Google Scholar] [CrossRef]
  66. Kumari, M.; Pudake, R.N.; Sing, V.; Breeding, P. Identification of microsatellite markers associated with staygreen trait in wheat RILs. Indian J. Genet. 2012, 72, 415–420. [Google Scholar]
  67. Paliwal, R.; Röder, M.S.; Kumar, U.; Srivastava, J.; Joshi, A.K.; Genetics, A. QTL mapping of terminal heat tolerance in hexaploid wheat (Triticum aestivum L.). J. Theor. 2012, 125, 561–575. [Google Scholar]
  68. Huang, S.S.; Sun, L.Q.; Hu, X.; Wang, Y.H.; Zhang, Y.J.; Nevo, E.; Peng, J.; Sun, D. Associations of canopy leaf traits with SNP markers in durum wheat (Triticum turgidum L. durum (Desf.)). PLoS ONE 2018, 13, e0206226. [Google Scholar] [CrossRef] [PubMed]
  69. Vijayalakshmi, K.; Fritz, A.K.; Paulsen, G.M.; Bai, G.; Pandravada, S.; Gill, B.S. Modeling and mapping QTL for senescence-related traits in winter wheat under high temperature. Mol. Breed. 2010, 26, 163–175. [Google Scholar] [CrossRef]
  70. Breseghello, F.; Sorrells, M.E. Association mapping of kernel size and milling quality in wheat (Triticum aestivum L.) cultivars. Genetics 2006, 172, 1165–1177. [Google Scholar] [CrossRef]
  71. Ma, J.F.; Liu, Y.; Zhang, P.P.; Chen, T.; Tian, T.; Wang, P.; Che, Z.; Shahinnia, F.; Yang, D. Identification of quantitative trait loci (QTL) and meta-QTL analysis for kernel size-related traits in wheat (Triticum aestivum L). BMC Plant Biol. 2022, 22, 607. [Google Scholar] [CrossRef]
  72. Zhang, J.; Lu, X.; Song, X.; Yan, J.; Song, T.; Dai, J.; Rocheford, T.; Li, J.S. Mapping quantitative trait loci for oil, starch, and protein concentrations in grain with high-oil maize by SSR markers. Euphytica 2008, 162, 335–344. [Google Scholar] [CrossRef]
  73. Kumari, M.; Pudake, R.; Singh, V.; Joshi, A.K. Association of staygreen trait with canopy temperature depression and yield traits under terminal heat stress in wheat (Triticum aestivum L.). Euphytica. 2013, 190, 87–97. [Google Scholar] [CrossRef]
  74. Talukder, S.K.; Babar, M.A.; Vijayalakshmi, K.; Poland, J.; Prasad, P.V.V.; Bowden, R.; Fritz, A. Mapping QTL for the traits associated with heat tolerance in wheat (Triticum aestivum L.). BMC Genom. 2014, 15, 97. [Google Scholar]
  75. Hussain, W.; Baenziger, P.; Belamkar, V.; Guttieri, M.J.; Venegas, J.P.; Easterly, A.; Sallam, A.; Poland, J. Genotyping-by-Sequencing Derived High-Density Linkage Map and its Application to QTL Mapping of Flag Leaf Traits in Bread Wheat. Sci. Rep. 2017, 7, 16394. [Google Scholar] [CrossRef] [PubMed]
  76. Xiao, Y.H.; Pan, Y.; Luo, L.H.; Deng, H.B.; Zhang, G.L.; Tang, W.B.; Chen LiYun, C.L. Quantitative Trait Loci Associated with Pollen Fertility under High Temperature Stress at Flowering Stage in Rice (Oryza sativa). Rice Sci. 2011, 18, 204–209. [Google Scholar] [CrossRef]
  77. Shirdelmoghanloo, H.; Taylor, J.D.; Lohraseb, I.; Rabie, H.; Brien, C.; Timmins, A.; Martin, P.; Mather, D.E.; Emebiri, L.; Collins, N.C. A QTL on the short arm of wheat (Triticum aestivum L.) chromosome 3B affects the stability of grain weight in plants exposed to a brief heat shock early in grain filling. BMC Plant Biol. 2016, 16, 100. [Google Scholar] [CrossRef] [PubMed]
  78. Kilasi, N.L.; Singh, J.; Vallejos, C.E.; Ye, C.; Jagadish, S.K.; Kusolwa, P.; Rathinasabapathi, B. Heat stress tolerance in rice (Oryza sativa L.): Identification of quantitative trait loci and candidate genes for seedling growth under heat stress. Front. Plant Sci. 2018, 9, 1578. [Google Scholar] [CrossRef] [PubMed]
  79. Park, J.R.; Yang, W.T.; Kim, D.H.; Kim, K.M. Identification of a novel gene, osbht, in response to high temperature tolerance at booting stage in rice. J. Mol. Sci. 2020, 21, 5862. [Google Scholar] [CrossRef]
  80. Al-Ashkar, I.; Alderfasi, A.; El-Hendawy, S.; Al-Suhaibani, N.; El-Kafafi, S.; Seleiman, M.F. Detecting Salt Tolerance in Doubled Haploid Wheat Lines. Agronomy 2019, 9, 211. [Google Scholar] [CrossRef]
  81. Sallam, M.; Ghazy, A.; Al-Doss, A.; Al-Ashkar, I. Combining Genetic and Phenotypic Analyses for Detecting Bread Wheat Genotypes of Drought Tolerance through Multivariate Analysis Techniques. Life 2024, 14, 183. [Google Scholar] [CrossRef]
  82. Grzesiak, S.; Hordyńska, N.; Szczyrek, P.; Grzesiak, M.T.; Noga, A.; Szechyńska-Hebda, M. Variation among wheat (Triticum easativum L.) genotypes in response to the drought stress: I–selection approaches. J. Plant Interact. 2019, 14, 30–44. [Google Scholar] [CrossRef]
  83. El-Hendawy, S.; Al-Suhaibani, N.; Al-Ashkar, I.; Alotaibi, M.; Tahir, M.U.; Solieman, T.; Hassan, W.M. Combining Genetic Analysis and Multivariate Modeling to Evaluate Spectral Reflectance Indices as Indirect Selection Tools in Wheat Breeding under Water Deficit Stress Conditions. Remote Sens. 2020, 12, 1480. [Google Scholar] [CrossRef]
  84. Al-Ashkar, I.; Sallam, M.; Al-Suhaibani, N.; Ibrahim, A.; Alsadon, A.; Al-Doss, A. Multiple Stresses of Wheat in the Detection of Traits and Genotypes of High-Performance and Stability for a Complex Interplay of Environment and Genotypes. Agronomy 2022, 12, 2252. [Google Scholar] [CrossRef]
  85. Abdolshahi, R.; Nazari, M.; Safarian, A.; Sadathossini, T.S.; Salarpour, M.; Amiri, H. Integrated selection criteria for drought tolerance in wheat (Triticum easativum L.) breeding programs using discriminant analysis. Field Crop. Res. 2015, 174, 20–29. [Google Scholar] [CrossRef]
  86. Chakraborty, K.; Monda, L.S.; Ray, S.; Samal, P.; Pradhan, B.; Chattopadhyay, K.; Kar, K.M.; Swain, P.; Sarkar, R. Tissue Tolerance Coupled With Ionic Discrimination Can Potentially Minimize the Energy Cost of Salinity Tolerance in Rice. Front. Plant Sci. 2020, 11, 265. [Google Scholar] [CrossRef]
  87. Saghai-Maroof, M.A.; Soliman, K.M.; Jorgensen, R.A.; Allard, R.W. Ribosomal DNA spacer-length polymorphisms in barley: Mendelian inheritance, chromosomal location, and population dynamics. Proc. Natl. Acad. Sci. USA 1984, 81, 8014–8018. [Google Scholar] [CrossRef]
  88. Lorenz, T.C. Polymerase chain reaction: Basic protocol plus troubleshooting and optimization strategies. J. Vis. Exp. 2012, 63, e3998. [Google Scholar]
  89. Aebi, H. [13] Catalase in vitro. In Methods in Enzymology; Elsevier: Amsterdam, The Netherlands, 1984; Volume 105, pp. 121–126. [Google Scholar]
  90. Chance, B.; Maehly, A.C. Preparation and assays of enzymes. Methods Enzymol. 1955, 2, 773–775. [Google Scholar]
  91. Duckworth, H.W.; Coleman, J.E. Physicochemical and kinetic properties of mushroom tyrosinase. J. Biol. Chem. 1970, 245, 1613–1625. [Google Scholar] [CrossRef] [PubMed]
  92. Kono, Y. Generation of Superoxide Radical during Autoxidation of Hydroxylamine and an Assay for Superoxide-Dismutase. Arch. Biochem. Biophys. 1978, 186, 189–195. [Google Scholar] [CrossRef] [PubMed]
  93. Pütter, J. Peroxidases. In Methods of Enzymatic Analysis; Elsevier: Amsterdam, The Netherlands, 1974; pp. 685–690. [Google Scholar]
  94. Bradford, M.M. A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. J. Anal. Biochem. 1976, 72, 248–254. [Google Scholar] [CrossRef]
  95. American Association of Cereal Chemists; Approved Methods Committee. Approved Methods of the American Association of Cereal Chemists; AACC: St. Paul, MN, USA, 2000. [Google Scholar]
  96. Donkor, E.F.; Adjei, R.R.; Amadu, B.; Boateng, A.S. Genetic variability, heritability and association among yield components and proximate composition of neglected and underutilized Bambara ground nut (Vigna subterranea (L) Verdc) accessions for varietal development in Ghana. Heliyon 2022, 8, e09691. [Google Scholar] [CrossRef]
  97. Lesk, C.; Rowhani, P.; Ramankutty, N. Influence of extreme weather disasters on global crop production. Nature 2016, 529, 84–87. [Google Scholar] [CrossRef]
  98. Liu, B.; Asseng, S.; Müller, C.; Ewert, F.; Elliott, J.; Lobell, D.B.; Martre, P.; Ruane, A.C.; Wallach, D.; Jones, J.W.; et al. Similar estimates of temperature impacts on global wheat yield by three independent methods. Nat. Clim. Chang. 2016, 6, 1130–1136. [Google Scholar] [CrossRef]
  99. De Costa, W. A review of the possible impacts of climate change on forests in the humid tropics. Natl. Sci. Found. 2011, 39, 281–302. [Google Scholar] [CrossRef]
  100. Asseng, S.; Foster, I.; Turner, N.C. The impact of temperature variability on wheat yields. Glob. Chang. Biol. 2011, 17, 997–1012. [Google Scholar] [CrossRef]
  101. Bennett, D.; Izanloo, A.; Reynolds, M.; Kuchel, H.; Langridge, P.; Schnurbusch, T. Genetic dissection of grain yield and physical grain quality in bread wheat (Triticum aestivum L.) under water-limited environments. Theor. Appl. Genet. 2012, 125, 255–271. [Google Scholar] [CrossRef]
  102. Yu, Q.; Li, L.; Luo, Q.; Eamus, D.; Xu, S.; Chen, C.; Wang, E.; Liu, J.; Nielsen, D.C. Year patterns of climate impact on wheat yields. Int. J. Climatol. 2014, 34, 518–528. [Google Scholar] [CrossRef]
  103. Lo, M.-H.; Wey, H.-W.; Im, E.-S.; Tang, L.; Anderson, R.G.; Wu, R.-J.; Chien, R.; Wei, J.; AghaKouchak, A.; Wada, Y. Intense agricultural irrigation induced contrasting precipitation changes in Saudi Arabia. Environ. Res. Lett. 2021, 16, 064049. [Google Scholar] [CrossRef]
  104. Rahman, M.M.; Akter, R.; Abdul Bari, J.B.; Hasan, M.A.; Rahman, M.S.; Abu Shoaib, S.; Shatnawi, Z.N.; Alshayeb, A.F.; Shalabi, F.I.; Rahman, A.; et al. Analysis of climate change impacts on the food system security of Saudi Arabia. Sustainability 2022, 14, 14482. [Google Scholar] [CrossRef]
  105. Motawei, M.; Al-Doss, A.; Moustafa, K. Environment. Genetic diversity among selected wheat lines differing in heat tolerance using molecular markers. J. Food Agric. 2007, 5, 180. [Google Scholar]
  106. Sinha, D.; Maurya, A.K.; Abdi, G.; Majeed, M.; Agarwal, R.; Mukherjee, R.; Ganguly, S.; Aziz, R.; Bhatia, M.; Majgaonkar, A.; et al. Integrated genomic selection for accelerating breeding programs of climate-smart cereals. Genes 2023, 14, 1484. [Google Scholar] [CrossRef]
  107. Sadat, S.; Saeid, K.A.; Bihamta, M.R.; Torabi, S.; Salekdeh, S.G.H.; Ayeneh, G.A.L. Marker assisted selection for heat tolerance in bread wheat. World Appl. Sci. J. 2013, 21, 1181–1189. [Google Scholar]
  108. Dhillon, G.S.; Das, N.; Kaur, S.; Srivastava, P.; Bains, N.S.; Chhuneja, P. Marker assisted mobilization of heat tolerance QTLs from Triticum durum-Aegilops speltoides introgression lines to hexaploid wheat. Indian J. Genet. 2021, 81, 186–198. [Google Scholar]
  109. Khan, A.; Ahmad, M.; Ahmed, M.; Gill, K.S.; Akram, Z. Association analysis for agronomic traits in wheat under terminal heat stress. Saudi J. Biol. Sci. 2021, 28, 7404–7415. [Google Scholar] [CrossRef]
  110. Gahtyari, N.C.; Jaiswal, J.P.; Sharma, D.; Talha, M.; Kumar, N.; Singh, N.K. Genetic analysis and marker association of physiological traits under rainfed and heat stress conditions in spring wheat (Triticum aestivum L.). Genetika 2022, 54, 1049–1068. [Google Scholar] [CrossRef]
  111. Al-Ashkar, I.; Al-Doss, A.; Ullah, N. Accelerating Crop Improvement through Speed Breeding. In Climate-Resilient Agriculture, Vol 1: Crop Responses and Agroecological Perspectives; Hasanuzzaman, M., Ed.; Springer International Publishing: Cham, Switzerland, 2023; pp. 821–847. [Google Scholar]
  112. De Leon, T.B.; Linscombe, S.; Gregorio, G.; Subudhi, P.K. Genetic variation in Southern USA rice genotypes for seedling salinity tolerance. Front. Plant Sci. 2015, 6, 374. [Google Scholar] [CrossRef]
  113. Junaid, M.B.; El-Hendawy, S.; Al-Ashkar, I.; Al-Suhaibani, N.; Alotaibi, M. Integrating Agro-Morpho-Physiological Traits and SSR Markers for Detecting the Salt Tolerance of Advanced Spring Wheat Lines under Field Conditions. Agriculture 2023, 13, 2135. [Google Scholar] [CrossRef]
  114. Sodini, S.M.; Kemper, K.E.; Wray, N.R.; Trzaskowski, M. Comparison of Genotypic and Phenotypic Correlations: Cheverud’s Conjecture in Humans. Genetics 2018, 209, 941–948. [Google Scholar] [CrossRef]
  115. Hill, C.B.; Li, C. Genetic improvement of heat stress tolerance in cereal crops. Agronomy 2022, 12, 1205. [Google Scholar] [CrossRef]
  116. Mason, R.E.; Mondal, S.; Beecher, F.W.; Pacheco, A.; Jampala, B.; Ibrahim, A.M.H.; Hays, D.B. QTL associated with heat susceptibility index in wheat (Triticum aestivum L.) under short-term reproductive stage heat stress. Euphytica 2010, 174, 423–436. [Google Scholar] [CrossRef]
  117. Ashraf, M.; Foolad, M.R. Roles of glycine betaine and proline in improving plant abiotic stress resistance. Environ. Exp. Bot. 2007, 59, 206–216. [Google Scholar] [CrossRef]
  118. Brisson, L.F.; Zelitch, I.; Havir, E.A. Manipulation of catalase levels produces altered photosynthesis in transgenic tobacco plants. J. Plant Physiol. 1998, 116, 259–269. [Google Scholar] [CrossRef]
  119. Feng, B.H.; Zhang, C.X.; Chen, T.T.; Zhang, X.F.; Tao, L.X.; Fu, G.F. Salicylic acid reverses pollen abortion of rice caused by heat stress. BMC Plant Biol. 2018, 18, 245. [Google Scholar] [CrossRef]
  120. Van Breusegem, F.; Dat, J.F. Reactive oxygen species in plant cell death. J. Plant Physiol. 2006, 141, 384–390. [Google Scholar] [CrossRef]
  121. Al-Ashkar, I.; Alderfasi, A.; Ben Romdhane, W.; Seleiman, M.; El-Said, R.; Al-Doss, A. Morphological and genetic diversity within salt tolerance detection in eighteen wheat genotypes. Plants 2020, 9, 287. [Google Scholar] [CrossRef]
  122. Al-Ashkar, I.; Sallam, M.; Almutairi, K.F.; Shady, M.; Ibrahim, A.; Alghamdi, S.S. Detection of high-performance wheat genotypes and genetic stability to determine complex interplay between genotypes and environments. Agronomy 2023, 13, 585. [Google Scholar] [CrossRef]
  123. Al-Ashkar, I.; Al-Suhaibani, N.; Abdella, K.; Sallam, M.; Alotaibi, M.; Seleiman, M.F. Combining genetic and multidimensional analyses to identify interpretive traits related to water shortage tolerance as an indirect selection tool for detecting genotypes of drought tolerance in wheat breeding. Plants 2021, 10, 931. [Google Scholar] [CrossRef]
  124. Khan, M.M.H.; Rafii, M.Y.; Ramlee, S.S.; Jusoh, M.; Al Mamun, M. Hereditary analysis and genotype× environment interaction effects on growth and yield components of Bambara groundnut (Vigna subterranea (L.) Verdc.) over multi-environments. Sci. Rep. 2022, 12, 15658. [Google Scholar] [CrossRef]
  125. Boakye-Peprah, B.; Ofori, K.; Asante, I.; Parkes, E. Genetic variability of three cassava traits across three locations in Ghana. Afr. J. Plant Sci. 2013, 7, 265–267. [Google Scholar] [CrossRef]
  126. Falconer, D.S. Introduction to Quantitative Genetics; Pearson Education India: Bengaluru, India, 1996. [Google Scholar]
  127. Obala, J.; Saxena, R.K.; Singh, V.K.; Vechalapu, S.; Das, R.; Rathore, A.; Sameer-Kumar, C.V.; Saxena, K.; Tongoona, P.; Sibiya, J.; et al. Genetic variation and relationships of total seed protein content with some agronomic traits in pigeon pea (‘Cajanus cajan’ (L.) Millsp.). Aust. J. Crop Sci. 2018, 12, 1859–1865. [Google Scholar] [CrossRef]
  128. Burton, G. Qualitative inheritance in grasses. In Proceedings of the 6th International Grassland Congress, State College, PA, USA, 17–23 August 1952; Volume 1, pp. 17–23. [Google Scholar]
  129. Al-Ashkar, I.; Sallam, M.; Ghazy, A.; Ibrahim, A.; Alotaibi, M.; Ullah, N.; AL-Doss, A. Agro-physiological indices and multidimensional analyses for detecting heat tolerance in wheat genotypes. Agronomy 2023, 13, 154. [Google Scholar] [CrossRef]
  130. Hosseini, S.J.; Sarvestani, Z.T.; Pirdashti, H. Responses of some rice genotypes to drought stress. Int. J. Agric. 2012, 2, 475–482. [Google Scholar]
  131. Khan, A.S.; Ashfaq, M.; Asad, M.A. A correlation and path coefficient analysis for some yield components in bread wheat. Asian J. Plant Sci. 2003, 2, 582–884. [Google Scholar]
  132. Ayer, D.; Sharma, A.; Ojha, B.; Paudel, A.; Dhakal, K. Correlation and path coefficient analysis in advanced wheat genotypes. J. Agric. 2017, 15, 1–12. [Google Scholar] [CrossRef]
  133. Novo, P.; Science, s. Cultivar and year effect on grain filling of winter barley. J. Plant Breed. 2001, 45, 45–58. [Google Scholar]
  134. Gunasekaran, K.; Sivakami, R.; Sabariappan, R.; Ponnaiah, G.; Nachimuthu, V.V.; Pandian, B.A. Assessment of genetic variability, correlation and path coefficient analysis for morphological and quality traits in rice (Oryza sativa L.). J. Agric. Sci. Dig.-Res. J. 2017, 37, 251–256. [Google Scholar] [CrossRef]
  135. Nitish, D.; Baranwal, D.; Kuma, S. Genetic evaluation of spring wheat (Triticum aestivum L.) genotypes for yield and spot blotch resistance in Eastern Gangetic Plains of India. J. Afr. J. Biotechnol. 2014, 13, 1867–1875. [Google Scholar] [CrossRef]
  136. Singh, D.; Singh, M.; Sharma, K. Correlation and path-coefficient analysis among flag leaf-area, yield and yield attributes in wheat (Triticum aestivum L.). J. Cereal Res. Commun. 1979, 7, 145–152. [Google Scholar]
  137. Bojarian, M.; Asadi-Gharneh, H.A.; Golabadi, M. Factor analysis, stepwise regression and path coefficient analyses of yield, yield-associated traits, and fruit quality in tomato. Int. J. Veg. Sci. 2019, 25, 542–553. [Google Scholar] [CrossRef]
  138. Del Moral, L.G.; Rharrabti, Y.; Villegas, D.; Royo, C. Evaluation of grain yield and its components in durum wheat under Mediterranean conditions: An ontogenic approach. J. Agron. J. 2003, 95, 266–274. [Google Scholar] [CrossRef]
  139. Arshad, M.; Ali, N.; Ghafoor, A. Character correlation and path coefficient in soybean Glycine max (L.) Merrill. J. Pak. J. Bot. 2006, 38, 121. [Google Scholar]
  140. Dewey, D.R.; Lu, K. A correlation and path-coefficient analysis of components of crested wheatgrass seed production. Agron. J. 1959, 51, 515–518. [Google Scholar] [CrossRef]
  141. Milligan, S.B.; Gravois, K.A.; Bischoff, K.P.; Martin, F.A. Crop Effects on Genetic-Relationships among Sugarcane Traits. Crop Sci. 1990, 30, 927–931. [Google Scholar] [CrossRef]
  142. Nasri, R.; Kashani, A.; Paknejad, F.; Vazan, S.; Barary, M.; Sciences, A.L. Correlation, path analysis and stepwise regression in yield and yield component in wheat (Triticum aestivum L.) under the temperate climate of Ilam province, Iran. Indian J. Fundam. 2014, 4, 188–198. [Google Scholar]
  143. Rameez Iftikhar, R.; IhsanKhaliq, I.K.; Muhammad Kashif, M.K.; Ahmad, M.; Smiullah, S. Study of morphological traits affecting grain yield in wheat (Triticum aestivum L.) under field stress condition. Middle-East. J. Sci. Res. 2012, 11, 19–23. [Google Scholar]
  144. Marenco, R.A.; Antezana-Vera, S.A.; Nascimento, H.C.S. Relationship between specific leaf area, leaf thickness, leaf water content and SPAD-502 readings in six Amazonian tree species. Photosynthetica 2009, 47, 184–190. [Google Scholar] [CrossRef]
  145. Yuan, S.; Li, Y.; Peng, S.B. Leaf Lateral Asymmetry in Morphological and Physiological Traits of Rice Plant. PLoS ONE 2015, 10, e0129832. [Google Scholar] [CrossRef]
  146. Gladun, I.V.; Karpov, E.A. Distribution of Assimilates from the Flag Leaf of Rice during the Reproductive Period of Development. Russ Plant Physiol. 1993, 40, 215–218. [Google Scholar]
  147. Briggs, K.G.; Aytenfisu, A. Relationships between morphological characters above the flag leaf node and grain yield in spring wheats. J. Crop Sci. 1980, 20, 350–354. [Google Scholar] [CrossRef]
  148. Murata, Y. Studies on the photosynthesis of rice plants and its culture significance. Agric. Sci. 1961, 9, 1–169. [Google Scholar]
  149. Cook, M.G.; Evans, L.T. Some Physiological-Aspects of the Domestication and Improvement of Rice (Oryza Spp.). Field Crop. Res. 1983, 6, 219–238. [Google Scholar] [CrossRef]
  150. Blanco, F.F.; Folegatti, M.V. Estimation of leaf area for greenhouse cucumber by linear measurements under salinity and grafting. Sci. Agric. 2005, 62, 305–309. [Google Scholar] [CrossRef]
  151. Fowler, C.; Rasmusson, D.C. Leaf area relationships and inheritance in barley. Crop Sci. 1969, 9, 729–731. [Google Scholar] [CrossRef]
  152. Jia, H.; Wan, H.; Yang, S.; Zhang, Z.; Kong, Z.; Xue, S.; Zhang, L.; Ma, Z. Genetic dissection of yield-related traits in a recombinant inbred line population created using a key breeding parent in China’s wheat breeding. J. Theor. 2013, 126, 2123–2139. [Google Scholar] [CrossRef]
  153. Wu, Q.H.; Chen, Y.X.; Fu, L.; Zhou, S.H.; Chen, J.J.; Zhao, X.J.; Zhang, D.; Zhang, D.; Ouyang, S.; Wang, Z.; et al. QTL mapping of flag leaf traits in common wheat using an integrated high-density SSR and SNP genetic linkage map. Euphytica 2016, 208, 337–351. [Google Scholar] [CrossRef]
  154. Liu, C.G.; Zhou, X.Q.; Chen, D.G.; Li, L.J.; Li, J.C.; Chen, Y.D. Natural Variation of Leaf Thickness and Its Association to Yield Traits in Rice. J. Integr. Agric. 2014, 13, 316–325. [Google Scholar] [CrossRef]
  155. Zhang, C.; Zheng, B.; He, Y. Improving Grain Yield via Promotion of Kernel Weight in High Yielding Winter Wheat Genotypes. Biology 2021, 11, 42. [Google Scholar] [CrossRef] [PubMed]
  156. Hussain Wani, S.; Brajendra Singh, N.; Haribhushan, A.; Iqbal Mir, J. Compatible solute engineering in plants for abiotic stress tolerance-role of glycine betaine. Curr. Genom. 2013, 14, 157–165. [Google Scholar] [CrossRef] [PubMed]
  157. Mohi-Ud-Din, M.; Hossain, M.A.; Rohman, M.M.; Uddin, M.N.; Haque, M.S.; Dessoky, E.S.; Alqurashi, M.; Aloufi, S. Assessment of genetic diversity of bread wheat genotypes for drought tolerance using canopy reflectance-based phenotyping and SSR marker-based genotyping. Sustainability 2022, 14, 9818. [Google Scholar] [CrossRef]
  158. El-Rawy, M.A.; Hassan, M.I. Assessment of genetic diversity in durum and bread wheat genotypes based on drought tolerance and SSR markers. Plant Breed. Biotechnol. 2021, 9, 89–103. [Google Scholar] [CrossRef]
  159. Zhang, Q.; Zhou, Z.Q.; Yang, G.P.; Xu, C.G.; Liu, K.D.; Saghai Maroof, M.A. Molecular marker heterozygosity and hybrid performance in indica and japonica rice. Theor. Appl. Genet. 1996, 93, 1218–1224. [Google Scholar] [CrossRef]
  160. Boppenmaier, J.; Melchinger, A.E.; Brunklaus-Jung, E.; Geiger, H.H.; Herrmann, R.G. Genetic Diversity for RFLPs in European Maize Inbreds: I. Relation to Performance of Flint × Dent Crosses for Forage Traits. Crop Sci. 1992, 32, 895–902. [Google Scholar] [CrossRef]
  161. Jinxiong, S.; Guangyuan, L.; Tingdong, F.; Guangsheng, Y. Relationships between genetic diversity and hybrid performance in Oilseed rape (Brassica napus). Zuo Wu Xue Bao 2002, 28, 622–627. [Google Scholar]
  162. Martin, J.; Talbert, L.; Lanning, S.; Blake, N. Hybrid performance in wheat as related to parental diversity. Crop Sci. 1995, 35, 104–108. [Google Scholar] [CrossRef]
Figure 1. Dendrogram showing the clustering of 13 genotypes based on SSR markers.
Figure 1. Dendrogram showing the clustering of 13 genotypes based on SSR markers.
Agronomy 14 01999 g001
Figure 2. Biplot for the first two principal components in the principal components analysis of 13 wheat genotypes. The genotypes started with C and H means the control and heat condition.
Figure 2. Biplot for the first two principal components in the principal components analysis of 13 wheat genotypes. The genotypes started with C and H means the control and heat condition.
Agronomy 14 01999 g002
Figure 3. Dendrogram showing the clustering of 13 genotypes based on Euclidean distance.
Figure 3. Dendrogram showing the clustering of 13 genotypes based on Euclidean distance.
Agronomy 14 01999 g003
Table 1. Selection of influential marker (independent variables) with six studies traits (dependent variables) for control, heat, and relative change based on SMLR analysis.
Table 1. Selection of influential marker (independent variables) with six studies traits (dependent variables) for control, heat, and relative change based on SMLR analysis.
TraitsTreatmentsMarkersR2 Par.R2 Com. p-Value *
GYControlwmc5270.7900.790<0.0001
Heatwmc3260.6060.6060.002
gwm3370.1800.7860.016
Relative changewmc3260.6710.6710.001
gwm3370.1330.8050.026
TKWControlwmc1540.4720.4720.010
Heatwmc3260.7640.764<0.0001
Relative changewmc3260.6720.6720.001
FLAHeatwmc3260.6990.6990.0001
Relative changewmc3260.8660.866<0.0001
wmc650.0490.9150.037
PHControlwmc650.3280.3280.041
Heatwmc3260.3590.359<0.0001
wmc540.3180.677<0.0001
wmc650.2050.882<0.0001
gwm3690.0830.9650.002
Relative changegwm3690.6030.6030.002
GFDControlwmc650.7360.7360.000
Heatwmc3260.8730.873<0.0001
wmc650.0480.921<0.0001
wmc540.0520.9720.000
wmc740.0170.9890.005
gwm3370.0050.9940.047
Relative changewmc3260.7460.746<0.0001
gwm3370.1250.8710.001
wmc5270.0620.9330.006
wmc540.0320.9650.028
CiControlwmc3260.4620.4620.011
Heatwmc3260.7640.764<0.0001
Relative changewmc3260.6560.6560.001
Coefficient partial determination (R2 Par.), cumulative coefficient determination (R2 Com.), * means p-value of coefficient partial determination. Abbreviation as described in Section 2.
Table 2. Analysis of 19 traits estimated in 13 wheat genotypes under heat stress and control.
Table 2. Analysis of 19 traits estimated in 13 wheat genotypes under heat stress and control.
SourceDFPPOCATPODSODGBPnCiFLAGFDGFR
H10.35 **103.99 **215.662 **4.54 **33.28 **1067.31 **171,829.81 **3924.31 **284.62 **308.07 **
Error A20.0020.0050.6680.0040.0030.53724.4536.8880.2050.059
G120.023 **0.470 **5.706 **0.031 **1.568 **6.901 **1821.177 **277.864 **69.78 **6.026 **
H*G120.009 **0.172 **4.637 **0.018 **0.611 **2.964 **501.600 **61.524 **26.10 **2.500 **
Error B480.0010.0120.130.0040.0210.59577.8379.5952.8960.327
Genetic Parameters
σ2G 0.0020.050.1780.0020.160.656219.92936.0577.2810.587
σ2e 0.0010.0020.0220.0010.0030.09912.9731.5990.4830.054
σ2Ph 0.0040.0780.9510.0050.2611.15303.52946.31111.6311.004
h2 % 59.39663.42518.7242.13461.04357.05272.45777.85862.60172.63
G.C.V. % 14.66612.6712.81511.20315.9338.6567.16114.8846.66.07
Ph.C.V. % 19.0315.90929.61917.25920.39211.468.41316.8688.3427.123
GA 0.0760.3660.3760.0630.6431.2626.00510.9154.3989.342
GG % 23.28420.78511.42214.98125.64313.46912.55727.05510.75710.65
SourceDFPHSLNGNSNSSTKWGYPCGI
Rep21.4232.01319.828217.7821.591.340.0270.74528.057
H12820.013 **29.538 **4078.154 **233,317.385 **108.513 **997.064 **92.966 **430.755 **1829.522 **
Error A217.3210.34641.734160.7310.3590.2370.0130.77952.474
G12233.902 **11.959 **207.238 **5597.440 **7.504 **48.428 **1.629 **7.021 **731.008 **
H*G1264.013 **0.733 ns76.916 *2051.051 **2.957 **20.011 **0.728 **3.685 **179.671 **
Error B488.1910.47133.13387.7010.8081.20.0150.72317.454
Genetic Parameters
σ2G 28.3151.87121.72591.0650.7584.7360.150.55691.889
σ2e 1.3650.0795.52214.6170.1350.20.0020.1212.909
σ2Ph 38.9841.99334.54932.9071.2518.0710.2711.17121.835
h2 % 72.63393.87262.88563.35760.59258.67955.2947.51775.421
G.C.V. % 6.07114.30212.525.045.4855.1755.9185.47112.42
Ph.C.V. % 7.12314.76215.7886.3327.0466.7567.9587.93614.301
GA 9.3422.737.61339.8641.3963.4340.5931.05917.149
GG % 10.65828.54520.4528.2648.7958.1679.0647.76822.22
* = significant at p ≤ 0.05, ** = significant at p ≤ 0.01, ns = insignificant. Abbreviation as described in Section 2.
Table 3. PCA of 13 wheat genotypes: eigenvalues proportion, and cumulative variance for the first three components of measured traits under heat.
Table 3. PCA of 13 wheat genotypes: eigenvalues proportion, and cumulative variance for the first three components of measured traits under heat.
PCA1PCA2PCA3
Eigenvalue12.1133.2041.217
Variability (%)63.75216.8616.404
Cumulative %63.75280.61387.018
Eigenvectors:
POD0.4120.4950.000
SOD0.6910.2440.000
PPO0.3370.3070.212
CAT0.7450.2080.007
GB0.3310.5610.012
Pn0.9740.0000.001
Ci0.9510.0030.001
FLA0.7390.1840.013
GFD0.4160.4150.016
GFR0.7130.0540.003
PH0.5070.0190.289
SL0.3740.2860.006
NS0.8750.0310.004
NSS0.7000.1800.001
NGS0.6230.0020.061
TKW0.7930.1490.003
GY0.9180.0210.000
Pc0.7850.0180.015
GI0.2310.0260.572
Values in bold indicate related traits, Abbreviation as described in Section 2.
Table 4. Stepwise regression analysis for grain yield, thousand-kernel weight, and flag leaf area (dependent index) with three yield-related traits (independent index).
Table 4. Stepwise regression analysis for grain yield, thousand-kernel weight, and flag leaf area (dependent index) with three yield-related traits (independent index).
Stepwise RegressionPath Coefficient
DependentSource Partitioning the CorrelationR2
Regression Coefficientp-ValueR2 Par.R2 Com.Direct EffectIndirect EffectCorrelation ValueDirect Effect
GYIntercept0.011
TKW0.5850.0040.8680.8680.4940.4370.9310.244
PH0.5300.0050.0660.9340.3430.4740.8170.118
Ci0.3160.0470.0250.9580.2590.5810.8400.067
Total direct effect 0.429
Total indirect effect 0.529
Total R2 0.958 0.958
Residual 0.205 0.205
TKWIntercept0.010
FLA0.481<0.0010.7800.780
Total direct effect
Total indirect effect
Total R2 0.780 0.780
Residua
FLAIntercept−0.117
GFD0.7940.0110.6990.6990.5110.3250.8360.261
Ci0.9510.0120.1460.8450.5020.3310.8330.252
Total direct effect 0.513
Total indirect effect 0.332
Total R2 0.845 0.845
Residual 0.394 0.394
Table 5. Posterior probability of membership in heat groupings through linear discriminant.
Table 5. Posterior probability of membership in heat groupings through linear discriminant.
Genotypes Classification Cross-Validation
PriorPosteriorMembership ProbabilitiesPosteriorMembership Probabilities
Pr(MT)Pr(S)Pr(T)MTST
16HTWYT-30SS0.0001.0000.000S0.0001.0000.000
DHL2MTMT1.0000.0000.000MT1.0000.0000.000
16HTWYT-20SS0.0001.0000.000S0.0001.0000.000
16HTWYT-38MTMT1.0000.0000.000MT1.0000.0000.000
16HTWYT-9SS0.0001.0000.000S0.0001.0000.000
KSU105TT0.0000.0001.000T0.0000.0001.000
16HTWYT-12SS0.0001.0000.000S0.0001.0000.000
Yecora RojoSS0.0001.0000.000S0.0001.0000.000
16HTWYT-22TT0.0000.0001.000T0.0000.0001.000
KSU115TT0.0000.0001.000T0.0000.0001.000
KlassicSS0.0001.0000.000S0.0001.0000.000
Line 47MTMT1.0000.0000.000MT1.0000.0000.000
KSU106TT0.0000.0001.000T0.0000.0001.000
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Sallam, M.; Al-Ashkar, I.; Al-Doss, A.; Al-Gaadi, K.A.; Zeyada, A.M.; Ghazy, A. Assessing Heat Stress Tolerance of Wheat Genotypes through Integrated Molecular and Physio-Biochemical Analyses. Agronomy 2024, 14, 1999. https://doi.org/10.3390/agronomy14091999

AMA Style

Sallam M, Al-Ashkar I, Al-Doss A, Al-Gaadi KA, Zeyada AM, Ghazy A. Assessing Heat Stress Tolerance of Wheat Genotypes through Integrated Molecular and Physio-Biochemical Analyses. Agronomy. 2024; 14(9):1999. https://doi.org/10.3390/agronomy14091999

Chicago/Turabian Style

Sallam, Mohammed, Ibrahim Al-Ashkar, Abdullah Al-Doss, Khalid A. Al-Gaadi, Ahmed M. Zeyada, and Abdelhalim Ghazy. 2024. "Assessing Heat Stress Tolerance of Wheat Genotypes through Integrated Molecular and Physio-Biochemical Analyses" Agronomy 14, no. 9: 1999. https://doi.org/10.3390/agronomy14091999

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