Next Article in Journal
The Combination of Whole-Brain Features and Local-Lesion Features in DSC-PWI May Improve Ischemic Stroke Outcome Prediction
Next Article in Special Issue
Adaptation of Some Quinoa Genotypes (Chenopodium quinoa Willd.), Grown in a Saharan Climate in Algeria
Previous Article in Journal
Surgical Treatment Options for Epiglottic Collapse in Adult Obstructive Sleep Apnoea: A Systematic Review
Previous Article in Special Issue
Evaluation of the Ecological Environment Affected by Cry1Ah1 in Poplar
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Molecular Breeding and Drought Tolerance in Chickpea

1
Department of Genetics & Plant Breeding, College of Agriculture, Rajmata Vijayaraje Scindia Krishi Vishwa Vidyalaya, Gwalior 474002, India
2
Department of Plant Molecular Biology & Biotechnology, College of Agriculture, Rajmata Vijayaraje Scindia Krishi Vishwa Vidyalaya, Gwalior 474002, India
3
Directorate of Research Services, Jawaharlal Nehru Agricultural University, Jabalpur 482004, India
*
Authors to whom correspondence should be addressed.
Life 2022, 12(11), 1846; https://doi.org/10.3390/life12111846
Submission received: 4 October 2022 / Revised: 5 November 2022 / Accepted: 7 November 2022 / Published: 11 November 2022
(This article belongs to the Special Issue Plant Biotic and Abiotic Stresses)

Abstract

:
Cicer arietinum L. is the third greatest widely planted imperative pulse crop worldwide, and it belongs to the Leguminosae family. Drought is the utmost common abiotic factor on plants, distressing their water status and limiting their growth and development. Chickpea genotypes have the natural ability to fight drought stress using certain strategies viz., escape, avoidance and tolerance. Assorted breeding methods, including hybridization, mutation, and marker-aided breeding, genome sequencing along with omics approaches, could be used to improve the chickpea germplasm lines(s) against drought stress. Root features, for instance depth and root biomass, have been recognized as the greatest beneficial morphological factors for managing terminal drought tolerance in the chickpea. Marker-aided selection, for example, is a genomics-assisted breeding (GAB) strategy that can considerably increase crop breeding accuracy and competence. These breeding technologies, notably marker-assisted breeding, omics, and plant physiology knowledge, underlined the importance of chickpea breeding and can be used in future crop improvement programmes to generate drought-tolerant cultivars(s).

1. Introduction

Chickpea is a diploid annual crop that is extremely self-pollinated [1]. After the faba bean and field pea, it is the world’s third most significant pules crop [2]. It is a popular cool-season legume crop with a 738-megabyte genome size [3]. With an annual production of 10.13 million tonnes from a land area of 9.44 million hectares and a productivity of 1073 kg ha−1, India is the greatest producer of chickpeas in the world [4]. Chickpeas are grown in 52 countries, together with Africa, Asia, Australia, and South Europe [5]. Mexico, Turkey, Canada, Iran, Australia, Tanzania, Ethiopia, Spain, and Burma are also notable producers of chickpea. Its seeds come in two varieties. The ‘desi’ chickpea is hardy in character, while the Kabuli chickpea has a delicate seed coat and appears to have evolved from the desi varieties [6,7]. In semi-arid zones, chickpea is cultivated in the form of a dry weather crop [8]; however, in cold climatic zones, it is grown as a rainfed crop [9,10]. In actuality, about 90% of the chickpea crop is cultivated in a rainfed environment [11,12,13]. Without irrigation, the crop is affected [14] at vegetative as well as reproductive phases. After illnesses, drought is the second most significant constraint to the yield of chickpea crop [15]. Drought has been reported as a factor of 40–50 percent yield reduction in chickpea [11,12,16,17,18].
The chickpea is also termed as the “poor man’s meat” [19], since it is important for supplying protein sources [20]. Nutritionists have also highlighted its importance due to high nutritional contents in it [21]. Chickpea is high in lysine and arginine [22], but low in methionine and cystine [23]. In general, the Kabuli type contains more protein than the desi kinds. It contains more calcium and phosphorus than most other pulse crops [24,25]. Chickpea seeds comprise 23% protein, 64% total carbohydrates (47% starch, 6% soluble sugar), 5% fat, 6% crude fibre, and 2% ash on average, as well as micronutrients, for example phosphorus, calcium, magnesium, iron, and zinc [26]. Recently, Singh et al. [21] also reported chickpea as good source of Fe and Zn. Consequently, Samineni et al. [27] examined the effects of drought stress on nutritional parameters of chickpea and observed significant differences in the nutritional contents due to stress.
Chickpea is mostly cultivated in the post-rainy season [28], using soil moisture that has been retained from the previous rainy season [29]. As a result, the crop is frequently subjected to severe heat and drought pressure [12,13,22]. Drought, among other abiotic factors, has a significant impact on chickpea output [30]. Drought and heat stress have been reported to have reduced chickpea yields by about 50% due to the damaging effects of the membrane and reduced photosynthesis [31].
The four climatic elements that are changing will have an impact on how much water plants consume [32]. These elements include rising CO2 concentrations and temperatures, more erratic precipitation, and changes in humidity. Due to the increased variability in precipitation during the growing season and more so in soils with low water holding capacity, these climate changes may result in an increase in the atmospheric water demand by crops and an increase in the potential for limitations in the availability of water in the soil. In the long run, breeding cultivars with high water use efficiency (WUE) is a more realistic and cost-effective strategy for raising yields in drought-prone locations. WUE promotes modest water absorption while maintaining elevated WUE, which is a crucial component of breeding programmes because of its yields in drought-prone areas. Any WUE is impacted by changes above the soil surface because they have an impact on the soil water balance by the evaporation and penetration of soil water. The majority of the chickpea crop is grown on residual moisture; however, additional irrigation can increase yields. At some sites in India, irrigation during the pre-flowering stage and at the beginning of the pod fill led to an increase in yield. Chickpeas’ reproductive cycle was prolonged by irrigation, which also increased plant biomass and increased the number of pods per plant.
The greatest sustained surface winds of tropical storms range from 39 to 73 mph, and they are fast rotating storm systems with an organized centre over warm tropical oceans. These storms have a wide range in size and can cause a variety of dangers for the impacted areas, including tornadoes, catastrophic winds, coastal floods, and inland flooding. The effects of tropical cyclones on drought have been extensively studied, but less research has been conducted on how smaller tropical storms affect the severity of drought. According to research, rainfall is not necessarily inversely correlated with the strength of a tropical cyclone; therefore, tropical storms can sometimes provide more rain than expected. The question of whether tropical storms can help to lessen and mitigate drought conditions is now being researched. Water deficit and surplus are related to drought and tropical storms (TS), respectively. When it comes to monitoring dryness, soil moisture is a crucial element of the hydrological cycle, since it reflects the water that TS rainfall has penetrated or stored. Soil moisture data can be used to determine whether TS can alleviate extremely severe drought situations [33]. The authors calculated the frequency of TS afflicted places in the US, including the ratio of droughts that TS exacerbated and alleviated, and the regions where TS have a significant impact on the offset of drought. Based on a high-resolution data set, the findings demonstrate extensive spatial information about the offset of drought conditions and offer potential guidance for future drought and TS mitigation.
Drought has a substantial influence on crop growth and photosynthesis, both of which are directly related to production [34,35]. Drought researchers must assess growth as well as physiological responses such as chlorophyll index [36], relative water content [37], membrane stability index, and biomass when determining the influence of drought on various crop metrics. The quantitative character of attributes and the prevalence of linkage between desired and undesired genes make developing drought-tolerant agricultural variants difficult [38]. Many experiments on the effects of drought on numerous chickpea features, such as root attributes, shoot biomass, and early maturity, have been conducted [39]. In this crop, various experiments have been performed successfully and published with specific conclusions on different aspects, such as morphological, physiological, biochemical, and molecular characteristics [40,41,42].
Advances genomics has made it possible to tag genes [43] associated to agronomic qualities, as well as the tolerance/resistance to abiotic and biotic challenges [44]. It is playing a significant role in the transfer of labelled genes through molecular breeding [45,46], quickly and accurately. In chickpeas, microsatellite and sequence-tagged microsatellite site markers have been found to be more beneficial [47,48]. As stress resistance/tolerance is governed by numerous genes [49], quantitative trait loci (QTL) mapping has proven to be effective in identifying and tagging the genes [50] involved for disease resistance/tolerance in plants. Foreground, recombinant and background selection are all examples of marker-assisted backcrossing. Linkage disequilibrium (LD) and association mapping are also determined using markers [51].
Next-generation sequencing (NGS) is a segment of revolutionary biology being a frontier area in crop science and produces correct data, with the results of significant throughput [52] and reduction in the need for fragment-cloning processes, which were the initial requirement for Sanger sequencing. NGS is used for the identification and mapping of mutations in targeted genotype [53,54]. Aside from whole genome sequencing (WGS), NGS also provides a platform for whole transcriptome shotgun sequencing, which is also termed as RNA sequencing (RNA-seq) [55,56] and whole-exome sequencing [57], which exhibits for functional variations [58], targeted or candidate gene sequencing [59].In the examination of large numbers of samples, RNA-seq enables a more precise and sensitive measurement of gene expression levels than microarrays [60].
Transcriptomics is the technology used to study the transcriptome of an organism [61]. Transcriptome is the complete set of genes [62] expressed under specific conditions by the genome of the targeted organism. MicroRNA (miRNA), transfer RNA (tRNA), messenger RNA (mRNA), ribosomal RNA (rRNA) and other non-coding RNA are all found in the transcriptome (ncRNA). Transcriptomics of chickpea [63] has provided insight into mechanisms of drought tolerance/avoidance, as well as pathogenesis-related and developmental processes [64]. Transcriptomics may undoubtedly have a greater impact on chickpea breeding in the future, including the use of microarrays.
Proteomics is the study of whole protein complement in a cell, tissue or organism in detail [65]. Mass spectrometry and protein microarrays can be used to analyse the proteome [66].
Role of various genes of plants under drought stress conditions have been recognized clearly [67]. Drought responsive mechanisms are activated in response to drought stress, which is a regular occurrence in plants. Morphological and structural changes [68], drought-resistant gene expression, hormonal and other biochemical changes are among these pathways [69]. Environmental stresses have the ability to change the developmental behaviour of plants. These alterations in plant growth and development [70,71] mostly resulted in lower yields [72]. We attempted to review the status and progress based on the existing literature on drought stress tests conducted in the chickpea.

2. Drought in Chickpea

Chickpea productivity has been found to be around 995 kg ha−1 on a global scale, which is quite low [73]. Drought, terminal heat [74], excessive salt, and cold are abiotic variables [75], whereas Ascochyta blight, Fusarium wilt, and Helicoverpa are biotic factors that have been recognised as key drivers of yield reduction in chickpea [76]. Drought stress was identified as a major cause in around 50% of chickpea output losses worldwide.
Several factors are responsible for complexity of drought stress (Table 1), including severity of drought, stage of crop, and duration of drought stress [77]. Two types of drought stresses, i.e., terminal and intermittent, have been reported with their impacts on crop plants [78]. During terminal drought, soil water availability diminishes over time, potentially leading to severe drought stress later in crop development. Intermittent drought is defined as a series of short episodes of insufficient rain or irrigation that occur at different times during the growing season [79]. Due to its limited cultivation on marginal terrain, chickpea is suffering from terminal drought stress. Intermittent and terminal drought stress is caused by breaks in rainfall combined with less moisture in terminal growth stages [80]. Apart from morpho-physiological factors various genes and proteins are also responsible for drought tolerance in chickpea crop (Table 2).
Table 1. Relevance of various physiological traits contributing to drought adaptation in chickpea.
Table 1. Relevance of various physiological traits contributing to drought adaptation in chickpea.
Physiological TraitsRelated withReferences
Early phenology (early
flowering, early podding)
Drought escape/conservative
water-use strategy
[81,82,83]
Crop growth rate High water harvest[47]
Shoot biomassHigh shoot biomass at maturity
contribute to a higher grain yield
under drought
[84]
Pod abortion and seed fillingHigh seed/grain yield could help in drought and heat stress tolerance[85]
Biomass partitioning Greater biomass partitioning to grain helps in drought and heat stress tolerance[46,47,86]
Pod number; high pod number Grain yield and contributes to heat, drought tolerance[87]
Pod production Number of pods/plants is more
affected at early stage than late stage under drought stress
[88]
Specific leaf area SLA has a positive effect on grain
yield at reproductive stage
[89]
Cell membrane stability Related to drought, heat, and cold
tolerance
[30,90,91,92]
Canopy temperature
depression
Cooler canopy contributes to
drought avoidance and has a positive association with seed yield under drought stress, and it also contributes to heat stress tolerance
[93,94,95]
Canopy conductance Associated to both heat and drought stress tolerance[96]
Carbon isotope
Discrimination
Transpiration efficiency [97]
Recycling of CO2 inside the podMaintain seed filling [98]
Antioxidants enzymes,
proline, anthocyanin
content, trehalose, sucrose, and nonreducing sugars
Increase in antioxidant enzymes,
proline, trehalose and anthocyanin content during vegetative stage causes drought and cold stress tolerance
[99]
Relative water content Increase in relative water content
causes drought stress tolerance
[100,101]
Chlorophyll content;
carotenoid content
Higher chlorophyll content and
carotenoid content helps in heat
stress tolerance
[55,96]
(Na+ and K+) ion uptake (Na+ and K+) ion uptake cause
drought tolerance
[102]
Chlorophyll a fluorescence
FO, FM, PSII, ETR, FV/FM
Enable preventing PSII
photochemistry from damage and helps in both drought and heat stress tolerance
[102,103]
Plant transpiration rate Low plant transpiration rate helps in conserving soil water[104,105]
Transpiration efficiency It decides ultimate yield [106,107]
Early vigour Associated to both heat and drought stress tolerance[108]
Pollen traits (pollen viability, fertility, and pollentube germination)High pollen viability and fertility
under heat stress are associated to heat stress tolerance
[109]
Abscisic acid (ABA)Under drought increase in ABA
causes closure of stomata, thus
reducing assimilate production that leads to the inhibition of seed set
[108]
Root architectural trait prolific root system, root branch, root density root depth, root area, and root volumeProlific root system is associated to grain yield[47]
Deep rooting helps in using
conserved soil moisture from subsoil
and helps in avoiding terminal
drought stress
[108]
Table 2. List of some genes conferring adaptation to drought and other abiotic stresses in chickpea.
Table 2. List of some genes conferring adaptation to drought and other abiotic stresses in chickpea.
TreatmentTraitsGeneReferences
DroughtAbiotic stress response CarERF116[110]
DroughtBiotic and abiotic stressesAquaporins gene family (CaAQPs) [111]
Drought Drought stress responseDEGs[112]
Drought, heat and cold stressProcess of plant developmentCarLEA4[113]
Drought and heat stressRoot traits, plat morphology, transpiration, and yield traits Marker–trait association [47]
Due to the overabundance of wheat in irrigated areas in India, chickpea growth is primarily limited to rainfed areas. Crops in rainfed areas are experiencing water shortages, particularly during the sowing and terminal growth periods.
Soil and plant management are important for minimizing water stress. For this purpose, various experiments have been conducted with the applications of different agents. Gypsum can enhance overall plant growth, since it is a moderately soluble source of the crucial plant nutrients, calcium and sulphur. Gypsum supplements can also enhance the physical and chemical characteristics of soils, hence lowering nutrient concentrations in surface water runoff and reducing soil erosion losses. The most often used addition for reclaiming sodic soil is gypsum, which can also be found in synthetic soils used in nursery, greenhouse, and landscaping applications. Gypsum can be used for a variety of purposes in agriculture and horticulture, which could be advantageous to users. There are currently no recognized standards that outline the broad best management practices for using gypsum in agricultural applications.
Drought tolerance is a complex phenomenon that involves defence mechanisms as well as stress-induced signal responses [114,115]. Drought stress triggers a number of physiological, biochemical, and molecular responses (Figure 1) that can be classified into six categories: drought escape [116], avoidance [117], tolerance [118], resistance [119], abandonment [120], and drought adaptation [12,121]. Some chickpea genotypes have been identified as drought sensitive [122] and others as drought tolerant [123,124]. Plant breeders apply different ways of selection and development of drought tolerant crop genotypes. Different strategies are important to protect plants from harmful effects of drought.
Drought escape is the capacity of the plant to complete its life cycle before experiencing a major water deficit. Drought escape causes early flowering and maturity, as well as better yield potential, allowing plants to finish reproduction before drought strikes [125]. Crop longevity is governed in part by genotype and in part by the environment, and it impacts the crop’s ability to withstand climatic conditions such as drought. To achieve large seed yields, it is necessary to match the plant growth time to soil moisture availability. The genotypes with early maturity have the capacity to escape the terminal drought stress, whereas the genotypes with late maturity generally needs well-watered environments. The length of the growing phase and yield potential are positively associated with each other. In this consequence, the development of shorter duration crop is important for the reliable management of drought stress in the chickpea. The timing of flowering is a key feature of a plant’s response to extreme drought and high temperatures [126]. Early maturity lets the crop circumvent the passé of stress, hence short duration cultivars may be produced to minimise production loss from terminal dryness. However, under ideal growth conditions, the yield is often associated with crop duration, and any decrease in crop interval underneath the optimal will tax yield [29,52].
Drought escape is a critical strategy for preventing chickpea crop from drought [12]. Water supply is harmonised with phenological development in drought escape. Early maturity aids in escaping terminal dryness and is a key feature in germplasm screening. However, growers are frequently incapable to reorganize for early planting owing to climatic factors [52].
Drought avoidance is explained as a plant’s aptitude to retain a high tissue water potential contempt in a lack of soil moisture [125]. Processes involved in the enhancement of water intake, its storage in plant cells, and limiting water loss are associated with drought avoidance. Other mechanisms, including deep rooting, increased level of hydraulic conductance, reduced level of epidermal conductance, radiation absorption and reduced leaf area have also been reported to be linked with drought avoidance in plants. Deep rooting promotes water intake, which is helpful in reducing water losses. In the chickpea, the stomata remains closed during the day to minimise water loss during drought, and as a result, the carbon assimilation is impeded, lowering production [126,127].
Root biomass plays major role in absorbing water [128], as it is advantageous even in the condition of less moisture in the soil. It means there is a linkage between the root system and water stress tolerance [129,130], thus, in the current scenario, breeders are focused in the development of cultivars with larger root systems [131]. From integrating large root features, cultivars have been developed by chickpea breeders with increased drought tolerance [43,132]. Because root size is governed by intrinsic genetic variables [133,134] and modified by multiple environmental signals, such as nutrition and moisture accessibility in the soil, it is a complicated feature [135]. During the vegetative growth stage, susceptible genotypes absorb more water than tolerant genotypes, whereas tolerant genotypes absorb more water during the reproductive stage [136]. The intake of water during the vegetative as well as reproductive stages of plants has a direct relation with seed yield [137]. The importance of roots, rather than just root growth, is determined by their temporal water intake [138]. The best method for screening the germplasm for water usage competence (WUE) is carbon isotope discernment (13C), and this method has also been adopted in the chickpea [139,140].
One of the important impacts of drought stress is stomatal closure. Drought stress reduces the stomatal conductance and transpiration rate. This declines the CO2 fixation and photosynthesis due to the reduction in the internal CO2 concentration of the leaf (Ci). All of these factors have their role on the reduction in yield due to the reduced rate of photosynthesis [141].
The reduced rate of photosynthesis is directly related to extreme drought stress, and it is a result of the decreased chlorophyll content. Because of the lowered chlorophyll content, continuous poor moisture availability reduces light collecting capacity, triggering the generation of reactive oxygen species due to excessive energy absorption [142]. This is also a cause of damaged photosynthetic machinery. The principal cause of chlorophyll depletion is reactive oxygen species [143]. Reduction in photosynthetic activities under drought stress, have been experimented in chickpea genotypes [55] and this reduction was found to be linked to reduced ATP synthesis [144,145]. The yield reduction in chickpea genotypes due to the flower and pod drop under heat and drought stress circumstances was also noticed [146,147].
The leaf surface is also an imperative characteristic of plants in relation to drought stress. As tiny leaf surfaces lose less water [148], waxy leaves have high water preservation potential. Waxy leaves have the ability of a reflectance of irradiation and the reduction of water loss. This helps in the reduction of leaf temperature and provides tolerance against drought condition. The preservation of water in leaves with a reduced leaf temperature are directly related to the drought tolerant behaviour of plants. Drought stress raises leaf temperature in a variable manner, as tolerance genotypes have lower leaf temperatures than sensitive genotypes [149]. One drought tolerant chickpea variety ‘Gokce’ has been developed by ICARDA through the gene pyramiding method, which can be survived under severe drought conditions. This variety possess some other important features, such as early maturity, resistance to Ascochyta blight, increased seed size, and suitability for mechanised harvesting [150].
The drought tolerance refers to a plant’s ability to maintain its metabolism in a water shortage [35] condition with low tissue water potential [58]. Two types of traits are responsible for the drought tolerance in plants, i.e., constitutive characters and acquired behaviours. The constitutive traits affect the yield at mild to moderate levels of drought stress, whereas the acquired traits affect the yield at severe levels of drought stress. Drought tolerance features are largely concerned with cellular structural protection against the effects of cellular dehydration. Due to a reduction in the plant tissue water content, dehydrins and late embryogenesis of abundant (LEA) proteins accumulate [151]. These proteins act as chaperones [152].
In recent years, the role of reactive oxygen species (ROS) in stress signalling has been widely researched and evaluated [153,154]. The extreme creation of ROS causes oxidative damage and, lastly, cell death [155]. The role of ROS as a signalling molecule or in the oxidative damage depends upon the equilibrium between production and the scavenging of them [156]. The scavenging of ROS under drought stress depends upon the action of antioxidants in the cell [157,158].
Pushpavalli et al. [159] emphasised the need of selecting chickpea genotypes that can withstand various shocks rather than simply one. High temperature stress, in addition to drought stress, is a new threat to chickpea production [33,160]. According to Kalra et al. [161], a temperature increase of 18 °C above a particular threshold causes a significant loss in chickpea output. Furthermore, it is predicted that a global temperature increase of 2–38 °C, along with erratic rainfall patterns, would pose a threat to chickpea yield. In agriculture, the yield is the most important parameter for crops, and a reduction in yield cannot be compromised at any level. There is a strong association between drought tolerance with yield in a crop [122,162]. This is because yield-related traits of crops have been found to be sensitive under drought stress [163].

3. Antioxidant Defence

Plants have multi-level systems of antioxidant defence [124] with main function to maintain homeostasis inside the cell. This system counteracts ROS and protects the cell from oxidative damage. In the absence of the sufficient quantity of an antioxidant to neutralize ROS, reactions such as biomolecule oxidation, lipid peroxidation, and protein damage, as well as nucleic acid (DNA, RNA) oxidation and apoptosis activation, may occur [90].
The antioxidant defence system has both enzymatic and non-enzymatic components. The enzymatic component involves superoxide dismutase, catalase, peroxidase, ascorbate peroxidase, and glutathione reductase. However, the non-enzymatic component involves cysteine, reduced glutathione, and ascorbic acid [164,165].
In plants, ascorbate peroxidase is an important antioxidant enzyme, and glutathione reductase is important for sustaining the reduced glutathione pool during stress [166]. In various plants, two glutathione reductase corresponding to deoxyribonucleic acids have been recognized, one producing cytosolic isoforms and the other encoding glutathione reductase proteins, which dually embattle chloroplasts and mitochondria [127]. Superoxide dismutase is a key enzyme that catalyses the detachment of two superoxide molecules into O2 and H2O2 [167]. The drought tolerance of a particular plant species can be linked to increased antioxidant enzyme activity [168].
Proline appears to play a variety of activities under stress situations as a multifunctional amino acid, including stabilizing proteins, membranes, and subcellular structures as well as defending cellular functioning by scavenging reactive oxygen species (ROS). The functional diversification of proline metabolism is more complicated as a result of the compartmentalization of proline production and degradation in the cytosol, chloroplast, and mitochondria. When the electron transport chain is saturated under stressful circumstances, the increased rate of proline production in the chloroplast can help to stabilize the redox balance and maintain cellular homeostasis by dissipating the excess reducing potential. Proline is one of the most widely dispersed suitable solutes and a key component of plant stress resistance that increases in plants under adverse environmental conditions. Proline serves as a superb osmolyte and also has important functions as a metal chelator, antioxidant defence molecule, and stress signalling molecule. By regulating mitochondrial activity, affecting cell growth, inducing certain gene expression, and stabilising membranes, it promotes stress tolerance by reducing electrolyte leakage, bringing ROS concentrations back into normal levels, and promoting stress recovery.
One of the elements driving drought resistance in the chickpea is proline build up in different plant sections due to the increased activity of proline synthesising enzymes [169]. Drought tolerant genotypes of chickpea had higher proline contents than sensitive genotypes [130]. Earlier, an increase in the leaf proline concentration under water-deprived conditions indicates an efficient osmotic regulating system in the chickpea. To modify the osmotic potential, proline, glycine betaine, and soluble carbohydrates are accumulated in response to drought stress [170].

4. Plant Growth Regulators

Plant growth regulators, basically known as phytohormones, can be administered externally or synthesised inside the plant [171]. Auxins, cytokinins, gibberellins, ethylene, and abscisic acid have all been referred to as plant growth regulators. However, in recent studies, brassinosteroids (BRs) and various compounds of jasmonic acid, cytokinin, salicylic acid, strigolactones, and some peptides have been identified as plant hormones. The concentrations of Auxins, gibberellins, and cytokinin are negatively related to drought, but abscisic acid and ethylene have a positive association [172]. Drought stress inhibits the formation of endogenous auxins, which is frequently accompanied by a rise in the levels of abscisic acid and ethylene [173].
Abscisic acid and cytokinin are thought to play opposing functions in drought stress. Under water stress, an upsurge in abscisic acid and a diminution in cytokinin levels favour stomatal closure and minimise water loss by transpiration [174]. Abscisic acid affects the relative growth rates of different plant parts, such as the root-to-shoot dry weight ratio, leaf area development inhibition, and the generation of prolific and deeper roots. It can influence the rate of transpiration by closing the stomata, and it may be implicated in the machinery providing drought tolerance in plants. Under drought conditions, ABA formation inhibits the lateral root growth [175].
Ethylene is a growth inhibitory hormone that acts as a part in both inhibiting and stimulating growth in response to environmental factors. The plants can maximise growth and resist abiotic challenges such as drought to avoid this adversity, and this response also requires ethylene synthesis [134].
Plant growth and development are known to be affected by polyamines. There has been an increasing interest in the role of polyamines in the plant defence against environmental stressors, and substantial research energies have been conducted in the last twenty years. The overexpression of the apple spermidine synthase gene, for example, results in high amounts of spermidine synthase, which enhances abiotic stress resistance, for instance, drought tolerance [176].

5. Role of Conventional Breeding

Breeders commonly employ traditional breeding techniques such as introduction, selection, hybridization, and mutation [106]. Hybridization is used to blend the desired characteristics from different parents into a single cultivar [85]. Any hybridization program’s success hinges on the selection of proper parents. Single, multiple, and three-way crosses [107] have all been employed for the hybridisations in the chickpea crop [177]. Among the different branches of breeding technology, the mutation breeding has been found as a powerful strategy [178] for the creation of genetic variability in crops [179]. It is considered under advanced breeding technologies [180]. According to Kumar et al. [181], fifteen chickpea varieties have been developed through mutation breeding and most of them are under the cultivation chain. The first chickpea variety developed in the year 1984 through mutation breeding in India was Kiran (RSG-2), which was the mutant form of RSG-10. This variety possess higher numbers of pods, early maturity, high yield, and tolerance to salinity stress [182].
The evaluation of different genotypes of a plant species in response to the drought controlled condition is needed [183]. In different studies, the phenopsis has been used as a non-automated control-guided drought screening method [145,184,185] to examine the performance of several Arabidopsis ecotypes [186]. In this regard, the ERECTA gene [147] responsible for the growth and development of the plant as well as the stomatal development, the ESKIMO1 gene [187], governs the plant water relation in Arabidopsis, which have been examined well. Similarly, some biosynthesis genes in the chickpea [188] have been well studied using controlled drought.
New cultivars, landraces, wild relatives, or a new crop species for the region could all be introduced. By using this technique, it is feasible to find a desirable genotype with a higher yield and better environmental tolerance, while also increasing the genetic variety. Through the international exchange of the germplasm and the inclusion of the crop of wild relatives and landraces, significant progress has been made in recent decades in enhancing the genetic diversity of the cultivated chickpea. Landraces are an important resource of novel genes in crop breeding. Landraces may possess genes for resistance against various biotic as well as abiotic stresses. For the identification of drought tolerant chickpea landraces, a field study was conducted by Kumar et al. [189]. The experiment included 37 chickpea landraces collected from ICARDA. Based on various morphological as well as physiological parameters, two landraces viz.,IG5856 (Jordan) and IG5904 (Iraq), were identified as drought tolerant.

6. Role of Molecular Breeding

Genetic Diversity

Complex abiotic stress such as drought requires a large group of genetic resources [190] to study the genetics of these stresses authentically. To fulfil the objective of molecular breeding in crop improvement, it is important to characterize the plant genetic resources. In molecular breeding, the characterization of plant genetic resources depends on the availability of DNA-based markers [191]. The molecular markers may be hybridization-based or PCR-based, depending on the technique of the detection of nucleotide variation [192]. The restriction fragment length of polymorphism (RFLP)is one of the molecular markers based on hybridization. Random-amplified polymorphic DNA (RAPD), amplified fragment length polymorphism (AFLP), simple sequence repeats (SSR) or microsatellite, sequence-tagged sites (STS), and cleaved amplified polymorphic sequence (CAPS) are all PCR-dependent molecular markers [158]. In the advanced category, single nucleotide polymorphism (SNP), single feature polymorphism (SFP), and diversity array technology (DArT) have been included [193]. In the field of crop improvement, molecular markers are categorized into dominant and co-dominant. The multi-locus markers (RAPD, ISSR, AFLP, etc.) come under the dominant category, while the single locus markers (SSR, STS, etc.) come under the co-dominant. This categorization of markers is basically based on their efficiency to discriminate homozygous and heterozygous genotypes. Dominant markers cannot differentiate homozygous and heterozygous genotypes. However, co-dominant markers have the ability to differentiate them.
In comparison to molecular markers, i.e., the hybridisation-based RFLP, the PCR-based RAPD and ISSR, SSR, and biochemical markers, i.e., isozyme, have a low polymorphic ability, which could be related to the decreased polymorphism in structural genes in the chickpea genome [194].
In some of the previous studies, the RAPD markers were utilised to detect genetic relationships amongst Cicer species [195]. The non-reproducibility nature of this dominant character is the main reason of the limited applicability of it [196]. However, the sequence characterised the amplified region (SCAR) markers developed with the use of RAPD markers, which are more suitable for the detection of the desired gene in crops, including the chickpea [197]. Amplification fragment length polymorphisms (AFLPs) were also reported to be uncommon in Cicer arietinum [198]. However, the availability of some reports on use of these markers in chickpea for diversity analysis [199] and the screening of abiotic as well as biotic resistant genotypes [200] proved their importance.
Microsatellite and STMS (sequence-tagged microsatellite site) markers are numerous, scattered throughout the genome, and highly polymorphic [198]. STMS raises the likelihood of finding polymorphism by a factor of ten. As a result, any genetic enhancement initiative should begin with a study of genetic variability. These markers become an important part of molecular breeding in the chickpea [201].
More than 3000 microsatellites [202,203], 15,000 DArT arrays [204,205], and SNPs [206] markers have been developed in the last few years in the chickpea. Because of a few of the specific characteristics, including co-dominance, abundance, repeatability, higher polymorphism and large genome coverage, SSR markers have proven their efficiency in the field of molecular breeding [207]. Subsequently, the data obtained after the use of SSR markers for molecular characterization or fingerprinting can be used to determine the genotypic identity of an individual. The applications of ISSR markers for the genotypic identification of chickpea genotypes in association with the seed germination and flowering time reported recently by Yadav et al. [208]. In this sequence, SNPs/InDels were also used recently for the gene identification and analysis in chickpea [209]. Basu et al. [210] identified SNPs linked with seed yield and Rajkumar et al. [211] identified SNPs linked with seed size and seed weight in the chickpea.

7. QTLs and Their Relevance with Drought Tolerance in Chickpea

A crucial requirement for identifying and integrating genes in linkage maps for marker–aided selection (MAS) is the knowledge of the agronomic trait inheritance [212]. The marker-assisted selection [54] and mapping of QTL (Quantitative Trait Loci) have been proposed to improve chickpea productivity [213]. Linkage map construction [214] and attribute mapping [215] were both conducted with available markers in the chickpea. Many research groups have focused their studies on abiotic stresses [216,217]. After completion of the sequencing of the desi and kabuli chickpea genomes [218], a genome-wide physical map was also generated. Furthermore, QTL studies have also been carried out in the chickpea (Table 3) to better understand the genetics of drought tolerance [217,219] and salt tolerance [220]. Varshney et al. [221] identified ‘QTL-hotspot’ regions that contain QTLs for a number of drought-related characteristics in the chickpea. They also reported the linkage between QTL hotspots and SSR markers. In the marker-assisted selection, the chickpea genotype ICC-4958 is used as a control for root studies under drought condition due to the large root character. This character makes this genotype suitable to use as a parent for transferring drought tolerance QTL-hotspot regions into the desired genotype. Recently, Muriuki et al. [222] also evaluated the root traits of some chickpea genotypes under drought stress and found that some of the desi genotypes (ICC4958, ICCV 00108, ICCV 92944 and ICCV 92318) performed fine.
Table 3. List of QTLs identified for drought tolerance in chickpea.
Table 3. List of QTLs identified for drought tolerance in chickpea.
Mapping
Approach
Numbers of
QTLs
Markers UsedStatistical Method UsedReferences
Biparental15 QTLs SSR [151]
Biparental93 QTLs SSR Composite interval
mapping-epistatic
mapping (ICIM-EPI)
[213]
Biparental and backcrossQTL-hotspot SSR, AFLP [13]
BiparentalQTL-hotspotSSR Composite interval
Mapping
[214]
GWAS 312 significant
model MTAs
DArT,
SNP
Mixed linear[58]
Biparental164 main-effect
QTLs
SNP,
CAPS
Composite interval
mapping
[215]
BiparentalQTL-hotspot_
a(15genes)
SNPICIM-ADD mapping
method
[216]
Biparental3 candidates
Genes
SNP [217]
Biparental 12 QTLs SNP [218]
Biparental21 QTLsSNPComposite interval
mapping
[223]
GWASSeveral MTAsSNP [224,225]
The advanced genomics involves a genome-wide association study (GWAS) that helps researchers in the screening of a wide range of genotypes [226,227] with different phenotypic or agronomic characters, and it also helps in the identification of the variability present among them. GWAS also helps in the identification of the association between the marker and a specific trait of interest [28,228]. The majority of these connotation investigations used either GWAS or candidate gene sequencing. Recently, the GWAS-based association mapping for the drought tolerance in the chickpea and for salinity tolerance have been performed. Apart from these, other examples of the association mapping in the chickpea are also available, i.e., for iron and zinc concentration in seeds [229] and Fusarium wilt resistance [230,231].
In some of the earlier studies, a combined analysis of the GWAS and sequencing of candidate gene [223] has been found to be more suitable in crop improvement. The GWAS study on two sets of chickpea genotypes with a different degree of their response in drought conditions helped in the discrimination of these genotypes on the basis of the single nucleotide polymorphisms generated.
Scientific efforts made on the improvement of the chickpea crop made it possible to generate not only a bi-parental plant population but also multi-parent populations [224]. The need of a multi-parental population was due to issues such as narrow genetic variability and limited efficiency of the bi-parental population [225] during the multiple trait analysis [232]. Multi-parent advanced generation inter-cross (MAGIC) populations for the chickpea are being established [233] to create diverse patterns of recombination [234]. The purpose of creating multi-parent populations is to advance the precision of QTL mapping [235] and discover specific loci regulating to the trait of interest [236]. ICRISAT and ICARDA played a major role in the development of a few MAGIC populations in the chickpea. One example of the MAGIC population developed at ICRISAT is the results of crossing eight varieties and advance breeding lines (ICC 4958, ICCV 10, JAKI 9218, JG 11, JG 130, JG 16, ICCV 97105, and ICCV 00108) with eight different founder parents [6,237]. Similar to this one, the MAGIC population developed at ICARDA was the result of crossing 12 different parents [238]. These plant populations accelerate the detection, isolation, and transfer of critical candidate genes to help in the development of chickpea varieties with superior agronomic traits [237]. One more approach (target-induced local lesions in genome -TILLING) [239] was adopted in the validation of the drought responsive gene in chickpea [58]. The selection and further use of agronomically superior genotypes of chickpea for the development of new varieties are the basic objectives of breeders involved in the chickpea crop improvement [240].

8. Attempts to Develop Drought Tolerant Varieties

One of the drought tolerant high yielding Ethiopian chickpea varieties,‘Geletu’, was developed and released in the year 2019 through the marker-assisted back-crossing after multi-location trials. During the development of this variety, the‘QTL-hotspot’ linked to drought tolerance was introgressed into an Indian chickpea cultivar JG11 from ICC4958 (gldc.cgiar.org). Recurrent selection is a crucial breeding technique used to increase crop plant populations. It is a productive method used in plant breeding to enhance the quantitative traits through repeated crossing and selection. Among the genomics-assisted selection methods, the marker-assisted back crossing has been found better for the introgression of the targeted region of the genome into a desired genotype. Consequently, the introgression of the ‘QTL hotspot’ region for the development of the drought-tolerant chickpea through molecular breeding has been found effective. Similar to this, numerous drought tolerance characters were introgressed into three elite Indian chickpea varieties: Pusa 372, Pusa 362, and DCP 92-3from ICC 4958. Recently, drought tolerant root traits have been introgressed into Kenyan chickpea varieties using the marker-assisted backcrossing approach [241].
Initially, the Pusa 372, chickpea variety was released as a drought tolerant variety for cultivation in the central, north-east, and north-west plains zones. However, under drought conditions, this variety’s output has decreased in recent years. To enhance drought tolerance in this variety, the MABC approach was adopted to introgress the ‘QTL-hotspot’ region from ICC 4958 into ‘Pusa 372′. Recently, the Pusa 372 was released with improved drought tolerance under the name ‘Pusa 10216′. This improved chickpea variety is the example of the first enhanced drought tolerant chickpea variety developed through the MABC approach. ICRISAT in collaboration with other research institutes in India is in the process of the development and release of drought tolerant chickpea varieties, i.e., IPC L4-14 and BGM 4005.Both of these varieties were developed by transferring a ‘QTL-hotspot’ from ICC4958 into DCP92-3 and Pusa362, respectively (https://www.icrisat.org/new-climate-resilient-disease-resistant-chickpea-varieties-coming-farmers-way/, accessed on 20 October 2022).

9. Whole-Genome Re-Sequencing

Whole-genome sequencing is the most thorough NGS technology [242], allowing for the complete genome sequencing and identification of variations in both exonic and non-coding areas, as well as the structural variant detection. Due to a paucity of genetic knowledge, the chickpea was formerly referred to as an orphan crop well adopted to suboptimal growing environments [243] However, researchers published the first draft genomes of the desi and kabuli chickpeas in 2013. The chickpea genome sequencing was based on advances in high-throughput sequencing and next-generation approaches. The BAC end genetic map and DArT markers were used to offer information on SSR and SNP molecular markers [244]. Both the kabuli and desi chickpea genomes have been updated, as well as a comparative examination of the two varieties. The QTLs associated to drought tolerance were reported by Jaganathan et al. [243]. The drought-responsive genomic areas were identified and employed in breeding approaches such as the marker-assisted gene interrogation and genetic gain to improve production in harsh climatic circumstances.
After publication of the draft genome sequence of the chickpea, the sequencing-based technique for the improvement of this crop has open multiple windows [244]. Furthermore, re-sequencing of a large number of chickpea assents collected from 45 nations enabled the identification of various candidate genes with their associations to a large number of agronomical characters [245]. The results of these experiments revealed the origin and migration routes of chickpea in the world. Re-sequencing data helped in the identification of 50,590 SNPs, and this data was used to develop the ‘Axiom®CicerSNP Array’ [246]. This SNP platform is being employed in the character mapping and identification of QTLs. Recently, Rajkumar et al. [247] reported the re-sequencing of large and small seed chickpea genotypes and 266 SNPs associated with seed size and seed weight. The findings of the study may help in the selection and categorization of chickpea genotypes on the basis of the size and weight of their seeds.
Next generation sequencing technology [248] has made possible the development of new markers for the improvement of the chickpea [249]. One of the important concepts, ‘The 3000 Chickpea Genome Sequencing Initiative’ [250], is an important step in the field of chickpea improvement. This initiative is helpful in the identification of variations in genomic sequences (rare alleles, markers) and their role in the determination of various agronomic characteristics, including yield and resistance/tolerance, against biotic and abiotic stresses. A thorough map of variation in 3171 farmed and 195 wild accessions was recently published in a project to give publicly accessible tools for chickpea genomics research and breeding [250]. This study also demonstrated the variations among the cultivated and wild progenitors of chickpea.

10. Pangenome and Super-Pangenome

The availability of the whole genome sequences of multiple individuals makes it feasible to compare them for the identification of diversity among them [251]. This approach may be termed as a comparative genomics [252] analysis, which allows for the identification of bio-markers linked with taxonomic as well as morphological and functional characteristics [253]. In this sequence, the pangenome concept was arisen, which allows for the accurate and efficient comparison of the genomes of a wide range of individuals [254]. A recently proposed revolutionary approach known as super-pangenome, allows the construction of the pangenomes of many species within a specified genus [255]. These concepts facilitate the identification of novel variations among individuals from different sources. These advanced technologies have their importance in crop breeding due to their accuracy and efficiency. The construction of pangenomes has their advantages in the identification of signature genomic areas relevant to crop domestication and evolution. Chickpea landraces and varieties have been sequenced to build the pangenome. These pangenome may be coupled with phenotypic traits and alleles associated with various characteristics and may also help in the identification of abiotic stress tolerance [256]. Pangenomes also provide a platform for the accurate identification of target genes for genome editing using CRISPR-clustered, regularly interspaced, short palindromic repeat technology [257].

11. Omics Approaches

Complex genetic traits, including drought tolerance, need advanced tools for their dissection, along with crop improvement [258]. Multiple omics approaches have revolutionized the identification of genes [259] as well as the metabolic database [260].Investigations have been carried out on transcriptomic analysis with the applications of the NGS technology in chickpea [261]. Multiple examples on transcriptomics’ evaluation in the chickpea are available as developing seeds [262], development and function [263], tissue specificity [264], and salinity tolerance [265] as well as root transcriptomics for drought tolerance. About 20,162 ESTs in the chickpea under salt and drought stress circumstances have been reported (Table 4). Recently, Kaashyap et al. [266] performed a comparative flower transcriptomic analysis to analyse the reproductive success under the salinity stress in the chickpea.
Table 4. Advanced technologies adopted to identify drought responsive differentially expressed genes/ESTs in chickpea.
Table 4. Advanced technologies adopted to identify drought responsive differentially expressed genes/ESTs in chickpea.
Differentially Expressed Genes/ESTsTechnique/Platform UsedReferences
1562 genes, 2592 genesIllumina HiSeq 3000[267]
1624 differentially expressed genesIllumina platform[103]
20,162 ESTs-[266]
53 ESTscDNA library [268]
3062 unigenesSuppression subtraction hybridization[258]
44,639 differentially expressed sequencesRoche/454 and Illumina/Solexa[269,270]
7532 unitags and 880 unitags SuperSAGE [267]
4053 and 1330 Illumina HiSeq 2000 platform[271,272]
261 (shoot) and 169 (root)Illumina TrueSeq RNA[273]
15,947 differentially expressed genesIllumina HiSeq 2000[274]
The RNA-Seq technique has also been used to analyse differential regulation of genes under drought stress in the chickpea [112]. Kumar et al. [113] analysed the gene expression of polyethylene glycol-stimulated drought stress in the chickpea, and thousands of differentially expressed genes (DEGs) were identified. Earlier, DEGs were detected in the kabuli chickpea under drought conditions [275]. Using RNA-Seq technique, it was used for the development of an inclusive C. arietinum Gene Expression Atlas (CaGEA) based on a drought tolerant ICC 4958 cultivar [269]. The findings of this study also validated the ‘QTL hotspot’ for drought tolerance in the chickpea.
The regulation of metabolic activities plays a major role in maintaining the osmotic potential of the cell under drought stress [276]. With the applications of the metabolomics method, many important metabolites were identified with a different regulation pattern during a drought [277] as well as in the salinity [278] in the chickpea. Similar to this, in an earlier investigation conducted on Arabidopsis thaliana, various genes were identified with their similar contributions under both salinity and drought stresses. The production of similar metabolites under both abiotic stresses indicates a common tolerance mechanism for drought as well as in the salinity in plants.
The proteomics’ analysis has also been performed in the chickpea for the identification of changes at a protein level under abiotic stresses. Earlier, a comparative proteomics analysis was conducted on one chickpea cultivar (JG-62), and novel dehydration-responsive proteins were detected [279]. In this sequence, Jaiswal et al. [280] reported the role of Sad1/UNC-84 in dehydration signalling. Recently, Vessal et al. [281] analysed the proteomic responses of drought sensitive and tolerant chickpea cultivars and identified changes in terms of the requirement of relative leaf water content for tolerant and susceptible cultivars. Drought responsive root proteins were also analysed recently by Gupta et al. [282].
Phenomics is an emerging tool in plant research, and is used to describe the use of genomics in phenotyping [283]. Phenomics studies for different phenotypic traits as well as seed yield have been conducted for the drought tolerance in chickpea [32]. The drought tolerance in chickpea is determined through phenotyping, and for this purpose, high-throughput screening technologies [32,44,284] have been adopted.

12. Role of Candidate Genes

A contender or candidate gene is thought to be linked to a specific disease or phenotypic character [285], and the biological function(s) of that has been derived either directly or indirectly from other investigations, including, for instance, the genome-wide association studies [286], the traditional map-based positional cloning technique, and the more recent next-generation sequencing (NGS) method [287]. Candidate gene studies are low-cost and rapid to conduct, and they focus on finding genes that have already been linked to the disease and hence have a prior knowledge of gene function [288]. Few of the important candidate genes detected in chickpea for the abiotic stress tolerance are Snf-1-related kinase (AKIN), DREB2A, dehydrin (DHN), CAP2, and Myb transcription factor (MYB) [289], Table 5. Despite the fact that multiple genes have been linked to drought resistance, the association study based on candidate gene sequencing has received little attention.
Table 5. List of various genes/transcription factors and their roles in response to drought and other abiotic stresses in chickpea.
Table 5. List of various genes/transcription factors and their roles in response to drought and other abiotic stresses in chickpea.
S.No.Gene/Transcription FactorFunctionReferences
1DREBDehydration responsive element binding proteins[290]
2Dehydrin (DHN)Response to water stress[291,292]
3STPKDrought stress[293]
4CADResponse to abiotic stress[294]
5AMADHWound healing, abiotic stress responsive[295,296]
6TCSAbiotic stresses tolerance[297]
7EREBPEthylene responsive[298]
8LEA GeneResponse to water stress[299]
9AKINPositive regulator of drought tolerance[300]
10Myb transcription factorStress[301]
11ASRAbscisic acid stress and ripening gene[302]
12SuSySucrose synthase[303]
13CAP2Promoter of DREB2A[297]
14ERECTATranspiration efficiency regulator[298]
15SPSSucrose phosphate synthase[300]
16CAMTASalinity and drought tolerance[304]
17CarNAC4Salt and Drought tolerance[305]
18CaNACDrought tolerance[306]
19CarERFDrought stress[109]
20CaSWEETAbiotic stress tolerance[307]

13. Transcription Factors and Their Role in Drought Tolerance in Chickpea

Transcription factors induce the cis-elements in the promoter provinces of different stress-responsive genes to accelerate the countenance of various downstream genes, which have their part in the stress tolerance [308] of plants. According to Riechmann et al. [309], nearly 1500 transcription factors have been reported in the Arabidopsis thaliana genome, which have their part in stress-responsive gene expression.

14. Dehydration Responsive Element Binding Proteins (DREBs)

DREBs (dehydration responsive element binding proteins) are key plant transcription factors [12]. They are responsible for the regulation of many stresses’ responsive genes. One of the transcription factors, AtDREB1a, was identified from Arabidopsis thaliana [310], with its role under abiotic stress. Recently, Das et al. [290] reported a better performance of AtDREB1a transgenic chickpea lines under water stress conditions.

15. Dehydrin (DHN)

Dehydrin (DHN) are stress-responsive proteins that are found when the temperature is low or when the body is dehydrated [311]. Protein dehydrin protects the embryo and seed tissues under water scarcity [312]. A better performance of transgenic plants overexpressing DHN than wild-type plants [313] have been identified. The role of the DHN gene in the Pusa1103 and Pusa362 genotypes of the chickpea has been found to be linked with the drought tolerance. Furthermore, in comparison to other genotypes, these genotypes were recognised as drought tolerant due to their better response [291].

16. Serine/Threonine Protein Kinase (STPK) Gene

Serine/threonine protein kinase (STPK), tyrosine protein kinase (TPK), and histidine protein kinase (HPK) are the three types of eukaryotic protein kinases [314]. In Arabidopsis, chickpea, and rice, the STPK family gene AtSnRK2.8 is reported with an enhanced degree of drought tolerance [293].

17. Cinnamyl Alcohol Dehydrogenase (CAD)

CAD is thought to be important in plant defence against a variety of biotic and abiotic stressors. When employing primers developed for the contig exhibiting match with the CAD gene of Arabidopsis thaliana, a homologue form of this gene was recovered from eight chickpea genotypes [294].

18. Ethylene-Responsive Element Binding Protein (EREBP) Gene

Ethylene-responsive element binding factors (ERFs) are a new type of transcription factor that are only found in plants. The ERF domain, a highly conserved DNA binding domain, is the protein family’s distinguishing trait. Primers for the chickpea were constructed using a contig sequence that was found to be identical to the Arabidopsis thaliana ethylene-responsive transcription factor. The amplification of eight chickpea genotypes yielded amplicons of around 400 bp. In plants, the AP2/EREBP genes have a variety of roles in developmental processes and stress responses [315].

19. Amino-Aldehyde Dehydrogenase (AMADH)

In some crops, an AMADH gene has been reported with its association to osmotic stress tolerance [316] by detoxifying hazardous aminoaldehydes; this gene has a function in physiological as well as metabolic responses under abiotic stresses [295]. On the basis of functional characterisation of the AMADH gene in Arabidopsis [317], the role of this gene should be examined in the chickpea.

20. ERECTA Gene

The ERECTA gene has a part in leaf organogenesis, lowering the density of the stomata on the leaf underside and thereby lowering evapotranspiration. It can also control the transpiration by the alteration of the leaf epidermal cell expansion, proliferation of mesophyll cells, and cell–cell interactions. The ERECTA gene has been demonstrated to regulate the growth and development of the organ and flower in Arabidopsis via encouraging cell proliferation [318]. Complementation tests on the wilting mutant Arabidopsis plants confirmed the involvement of the ERECTA gene to the water usage competence [319]. Pioneer Hi-Bred International, Inc. has patented the ZmERECTA genes from maize, which were implicated in the drought tolerance in crop plants.

21. Late-Embryogenesis Abundant (LEA) Proteins

The attainment of dehydration tolerance and the behaviour of plants to drought have been linked to late-embryogenesis abundant (LEA) proteins. Increased LEA and Dehydrin expression in genotypes during the vegetative, flowering, and podding stages could represent an adaptation to assist the plant survival by supplying the energy for growth and survival [29]. Leaf age inhibited arLEA4 expression, which changed during seed and pod development, including during germination. Drought, salt, heat, cold, ABA, IAA, GA3, and MeJA all significantly increased the CarLEA4 expression. CarLEA4 is a LEA assembly 4 protein that may participate in a variety of plant developmental processes as well as abiotic stress responses [299].

22. Myeloblastosis (MYB) Gene

Plants have a big transcription factor (TF) family, called the myeloblastosis (MYB) gene [320]. It plays a role in the secondary metabolism regulation, hormonal and climatic condition response, cell differentiation, and resistance to drought and other abiotic stimuli. Under drought stress, arrays of MYB-transcription factors are involved in the generation of epicuticular waxes [321]. These waxes seal the plant’s aerial component and reduce water loss through the leaf surfaces [322]. In an experiment, the root tissue of ICC 4958 (drought tolerant), ICC 1882 (drought sensitive), JG 11 (elite), and JG 11+ (introgression line) were employed to recognize the role of the 1R-MYB gene in the machinery of the drought tolerance in the chickpea. The findings of this experiment were suggested to conduct more experiments on this aspect in the chickpea. Recently, Caballo et al. [301] observed that CaRAX1/2a codes a MYB transcription factor that is exactly articulated in the meristem of chickpea. These results disclosed that the single flower gene (SFL) encodes for MYB, which works as a central factor responsible for the regulation of the numbers of flowers in chickpea inflorescence.

23. S-Adenosylmethionine Synthetase Gene

S-adenosylmethionine (SAM) is a precursor in the production of polyamines and ethylene [323]. In plants, the action of 1- aminocyclopropane-1-carboxylate (ACC) synthase and ACO (ACC oxidase) is responsible for ethylene biosynthesis, while the activity of SAM decarboxylase is responsible for spermidine and spermine production. The exogenous polyamine administration or overexpression of polyamine production genes has been demonstrated to improve abiotic stress tolerance. Primers were developed using a contig sequence that was comparable to the S-adenosylmethionine synthetase 1 (SAM1) gene of Arabidopsis thaliana for the isolation of the S-adenosylmethionine synthetase 1 gene homologue in the chickpea. The PCR amplification revealed amplicons of roughly 300 bp in eight chickpea genotypes [324].
Expression of SAM gene in pigeon pea (Cajanus cajan L.) was evaluated under drought, heavy metal (CdCl2), and cold stresses. The enhanced up-regulation of SAM gene in the leaves were recorded after three days [325].

24. Abscisic Acid Stress and Ripening Gene

Among many other genes, the abscisic acid stress and ripening (ASR) gene plays a critical role in controlling various plant stresses. The ASR gene has been reported in plants, and is induced by abscisic acid and different abiotic stresses during the process of fruit ripening [268]. Reports on ASR genes with their responses in different plant species under drought, salt, and cold stresses [326] confirm their role. Transgenic Arabidopsis demonstrated the over-expression of the ASR gene in response to drought and salt stresses [302]. Genotypes of rice also presented the association of the ASR gene expression [327]. Similarly, Cortés et al. [328] reported the potential significance of the ASR1 gene in the common bean. In a recent study conducted on the chickpea under drought stress, increased ASR gene expression was observed. The increased expression may have helped the drought-tolerant chickpea genotypes function better under stress. This hypothetical ASR protein could have boosted the activity of the ASR gene as a transcription factor mediating drought responses in chickpeas.

25. ABRE-Binding Protein (AREB)

Various genes that are activated by abscisic acid (ABA) have been discovered to be drought stress-inducible. Such ABA-regulated genes have conserved cis-elements in their promoter regions known as ABA responsive elements (ABREs), which use bZIP-type AREB/ABF transcription factors to regulate the gene expression. ABA and water stress upregulate the expression of the AREB/ABF gene. Expression of AREB gene under drought stress has been reported by Yoshida et al. [329] in Arabidopsis thanliana.

26. Sucrose Synthase (SuSy) Gene

Sucrose synthase (SuSy) is a crucial enzyme that hydrolyzes sucrose directly to provide substrates for plant metabolism. It is also used as a bio-marker for plant sink strength [330]. Plant sink strength improvement could contribute to increased plant growth and yield [331]). In an experiment, cultivars and treatments had a strong and positive association between the seed dry weight at maturity and peak sucrose synthase movement. Sucrose synthase is a decent physiological indication to employ in chickpea breeding for larger seeds.
Sucrose synthase activity has a major role in chickpea seed growth. The supremacy of the sink, as measured by the sucrose synthase movement most of the time, hinges upon genetic features of a genotype along with the accessibility of water obtainable at seed filling [332]. In both the large-seeded kabuli and the small-seeded desi varieties, the water shortage reduced the enzyme action and seed size, but the higher enzyme action in the large-seeded kabuli, pre-dominantly at the late seed filling stage, seemed to persuade a better remobilization of the integrates from the pod wall and seed coat. The cotyledons’ greater sucrose synthase action is taken into account. The strong association between sucrose synthase activity during rapid seed filling and final seed dry weight accumulation, and therefore seed size, advises that the sink strength is an important element of the seed size in chickpea. The tight link between the sucrose synthase activity during rapid seed filling and final seed dry weight build up, and thus seed size, implies that the sink strength is a key factor in the chickpea seed growth. Higher cotyledon sucrose synthase activity is vital in breeding for better seed size in chickpeas, regardless of the growth environment [303].

27. CAP2 Gene

The AP2 subgroup of proteins has two copies of the DNA-binding domain (BD), detached by an insertion province [333]. CAP2 is a C-Repeat binding factor (CBF) that muddles to the DRE/CRT (dehydration responsive element/C-repeat element)(CCGAC) found in the promoters of abiotic-stress responsive genes. Dehydration, excessive salinity, and exogenous ABA treatment all enhanced CAP2 gene expression. The incidence of roughly 60-amino-acid long AP2/ERF DNA-binding realms in these transcription regulators allows them to connect directly with GC-rich cis-acting elements (GCC box/C-repeat) in the promoter of their target genes. Ectopic expression of CAP2 in tobacco resulted in increased drought, salinity, and heat tolerance, as well as improved transgenic plant growth [334]. The enhanced accumulation of the CaZF transcript was caused by the transient expression of CAP2 in chickpea leaves. CAP2 activates the CaZF promoter through interacting with C-repeat elements (CRTs) in CaZF promoter, according to the gel mobility shift and transient promoter-reporter tests. The CAP2 protein interacts with the CaZF promoter in vivo, according to a chromatin immunoprecipitation (ChIP) test [335].

28. Sucrose Phosphate Synthase (SPS) Gene

The sucrose phosphate synthase (SPS) gene has an important function in the sucrose production in different plant species. It regulates sucrose metabolism in drought sensitive and tolerant genotypes [300]. Sucrose biosynthesis and sucrose degradation determines the level of sucrose in a genotype, and an optimum level of it is important for growth and development under environmental stress in plants [336]. Some reports are available on the unchanged or decreased level of the SPS activity in maize, potato, soybean, and some other crops, however, some reports advocate the increment in the SPS activity in rice, wheat crops, and Arabidopsis [336]. The significance of the SPS gene has been studied in chickpea under a low temperature by Sharma et al. [300].

29. Genome Editing Options

The genome editing approach, specially CRISPR-Cas9, has proved their efficiency in the development of climate resilient cultivars of different crops [337]. Two genes, namely RVE7 and 4CL, have been identified in the chickpea and their association with drought tolerance.The CRISPR/Cas9-mediated editing of the chickpea protoplast was reported for the first time by Badhan et al. [338], where they reported knock-outs of 4CL andRVE7 genes, which are associated with drought tolerance mechanisms. This report laid down a foundation for future genome editing options in the chickpea [339]. Genome editing approaches with the applications of CRISPR-Cas9 (Figure 2) may be helpful in the development of abiotic stress tolerance in chickpea genotypes including drought.

30. Conclusions

As previously stated, the changes in the plant shape and internal biochemical characteristics during drought stress have been extensively characterised in previous studies. Plant drought stress techniques can help us better use scientific means to improve plant tolerance to water shortage environments and increase crop yields, allowing us to play a larger role. As a result, by thoroughly examining and summarising the mechanisms of the chickpea plant response to drought, this study provides essential background knowledge and theoretical framework for selective breeding, cross breeding, and molecular breeding of the chickpea in the future. Drought tolerant land races/germplasm lines may be employed in classical as well as molecular breeding programmes to breed drought tolerant cultivars in future by using the available scientific data.

Author Contributions

Writing: R.A., S.T., R.K.Y. and N.T.; Editing: M.K.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Madurapperumage, A.; Tang, L.; Thavarajah, P.; Bridges, W.; Shipe, E.; Vandemark, G.; Thavarajah, D. Chickpea (Cicer arietinum L.) as a Source of Essential Fatty Acids–A Biofortification Approach. Front. Plant Sci. 2021, 12, 734980. [Google Scholar] [CrossRef] [PubMed]
  2. Singh, A.; Nath, O.; Singh, S.; Kumar, S.; Singh, I.K. Genome-wide identification of the MAPK gene family in chickpea and expression analysis during development and stress response. Plant Gene 2018, 13, 25–35. [Google Scholar] [CrossRef]
  3. Varshney, R.K.; Gaur, P.M.; Chamarthi, S.K.; Krishnamurthy, L.; Tripathi, S.; Kashiwagi, J.; Samineni, S.; Singh, V.K.; Thudi, M.; Jaganathan, D. Fast-track introgression of ‘QTL-hotspot’ for root traits and other drought tolerance traits in JG 11, an elite and leading variety of chickpea. Plant Genome 2013, 6, 3. [Google Scholar] [CrossRef] [Green Version]
  4. Directorate of Economics and Statistics. 2019. Available online: https://eands.dacnet.nic.in/ (accessed on 20 October 2022).
  5. Ganjeali, A.; Porsa, H.; Bagheri, A. Assessment of Iranian chickpea (Cicer arietinum L.) germplasms for drought tolerance. Agric. Water Manag. 2011, 98, 1477–1484. [Google Scholar] [CrossRef]
  6. Gaur, P.M.; Samineni, S.; Thudi, M.; Tripathi, S.; Sajja, S.B.; Jayalakshmi, V.; Mannur, D.M.; Vijayakumar, A.G.; Ganga Rao, N.V.; Ojiewo, C.; et al. Integrated breeding approaches for improving drought and heat adaptation in chickpea (Cicer arietinum L.). Plant Breed. 2019, 138, 389–400. [Google Scholar] [CrossRef] [Green Version]
  7. Gaur, R.; Verma, S.; Pradhan, S.; Ambreen, H.; Bhatia, S. A high-density SNP-based linkage map using genotyping-by-sequencing and its utilization for improved genome assembly of chickpea (Cicer arietinum L.). Funct. Integr. Genom. 2020, 20, 763–773. [Google Scholar] [CrossRef] [PubMed]
  8. Sahu, V.K.; Tiwari, S.; Gupta, N.; Tripathi, M.K.; Yasin, M. Evaluation of physiological and biochemical contents in desi and Kabuli chickpea. Legume Res. 2020, 45, 1197–1208. [Google Scholar] [CrossRef]
  9. Kumar, T.; Hamwieh, A.; Swain, N.; Sarker, A. Identification and morphological characterization of promising kabuli chickpea genotypes for short-season environment in central India. J. Genet. 2021, 100, 33. [Google Scholar] [CrossRef]
  10. Tahir, N.A.R.; Karim, H.F.H. Impact of magnetic application on the parameters related to growth of chickpea (Cicer arietinum L.). Jordan J. Biol. Sci. 2010, 3, 175–184. [Google Scholar]
  11. Mohammed, A.; Tana, T.; Singh, P.; Korecha, D.; Mollad, A. Management options for rainfed chickpea (Cicer arietinum L.) in northeast Ethiopia under climate change condition. Clim. Risk Manag. 2017, 16, 222–233. [Google Scholar] [CrossRef] [Green Version]
  12. Rani, A.; Devi, P.; Jha, U.C.; Sharma, K.D.; Siddique, K.H.M.; Nayyar, H. Developing Climate-Resilient Chickpea Involving Physiological and Molecular Approaches with a Focus on Temperature and Drought Stresses. Front. Plant Sci. 2020, 10, 1759. [Google Scholar] [CrossRef] [PubMed]
  13. Arif, A.; Parveen, N.; Waheed, M.Q.; Atif, R.M.; Waqar, I.; Shah, T.M. A Comparative Study for Assessing the Drought-Tolerance of Chickpea Under Varying Natural Growth Environments. Front. Plant Sci. 2021, 11, 607869. [Google Scholar] [CrossRef]
  14. Kumar, J.; Abbo, S. Genetics of flowering time in chickpea and its bearing on productivity in semiarid environments. Adv. Agron. 2001, 72, 107–138. [Google Scholar]
  15. Singh, K.B.; Malhotra, R.S.; Halila, M.H.; Knights, E.J.; Verma, M.M. Current status and future strategy in breeding chickpea for resistance to biotic and abiotic stresses. Euphytica 1994, 73, 137–149. [Google Scholar] [CrossRef]
  16. Millan, T.; Clarke, H.J.; Siddique, K.H.; Buhariwalla, H.K.; Gaur, P.M.; Kumar, J.; Gil, J.; Kahl, G.; Winter, P. Chickpea molecular breeding: New tools and concepts. Euphytica 2006, 147, 81–103. [Google Scholar] [CrossRef] [Green Version]
  17. Jameel, S.; Hameed, A.; Shah, T.M. Investigation of Distinctive Morpho-Physio and Biochemical Alterations in Desi Chickpea at Seedling Stage Under Irrigation, Heat, and Combined Stress. Front. Plant Sci. 2021, 12, 692745. [Google Scholar] [CrossRef]
  18. Sachdeva, S.; Bharadwaj, C.; Patil, B.S.; Pal, M.; Roorkiwal, M.; Varshney, R.K. Agronomic Performance of Chickpea Affected by Drought Stress at Different Growth Stages. Agronomy 2022, 12, 995. [Google Scholar] [CrossRef]
  19. Grewal, S.K.; Sharma, K.P.; Bharadwaj, R.D.; Hegde, V.; Tripathi, S.; Singh, S.; Jain, P.K.; Agrawal, P.K.; Mondal, B. Understanding genotypic variation and identification of promising genotypes for iron and zinc content in chickpea (Cicer arietinum L.). J. Food Compos. Anal. 2020, 88, 103458. [Google Scholar] [CrossRef]
  20. Sahu, V.K.; Tiwari, S.; Tripathi, M.K.; Gupta, N.; Tomar, R.S.; Yasin, M. Morpho-physiological and biochemical traits analysis for Fusarium wilt disease using gene-basedmarkers in desi and Kabuli genotypes of chickpea (Cicerarietinum L.). Indian J. Genet. 2020, 80, 163–172. [Google Scholar]
  21. Singh, S.; Babu, K.S.; Arora, A.; Panwar, R.K.; Verma, S.K. Genetic studies for biofortification traits in chickpea. J. Food Legumes 2021, 34, 17–20. [Google Scholar]
  22. Wallace, T.C.; Murray, R.; Zelman, K.M. The Nutritional Value and Health Benefits of Chickpeas and Hummus. Nutrients 2016, 8, 766. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Iqbal, A.; Ateeq, N.; Khalil, I.A.; Perveen, S.; Saleemullah, S. Physicochemical characteristics and amino acid profile of chickpea cultivars grown in Pakistan. J. FoodSer. 2006, 17, 94–101. [Google Scholar] [CrossRef]
  24. Singh, S.; Singh, D.; Rao, V.U.M. Seedling establishment of chickpea cultivars in varying sowing environments under field conditions. J. Agrometeorol. 2009, 11, 148–151. [Google Scholar] [CrossRef]
  25. Hirdyani, H. Nutritional composition of Chickpea (Cicer arietinum-L) and value added products. Indian J. Community Health Haryana J. Agron. 2014, 4, 116–118. [Google Scholar]
  26. Jukanti, A.K.; Gaur, P.M.; Gowda, C.L.L.; Chibbar, R.N. Chickpea: Nutritional properties and its benefits. Br. J. Nutr. 2012, 108, S11–S26. [Google Scholar] [CrossRef]
  27. Samineni, S.; Mahendrakar, M.D.; Shankar, N.; Hotti, A.; Chand, U.; Rathore, A.; Gaur, P.M. Impact of heat and drought stresses on grain nutrient content in chickpea: Genome-wide marker-trait associations for protein, Fe and Zn. Environ. Exp. Bot. 2022, 194, 104688. [Google Scholar] [CrossRef]
  28. Korbu, L.; Fikre, A.; Tesfaye, K.; Funga, A.; Bekele, D.; Ojiewo, C.O. Response of chickpea to varying moisture stress conditions in Ethiopia. Agrosyst. Geosci. Environ. 2022, 5, e20234. [Google Scholar] [CrossRef]
  29. Marjani, A.; Farsi, M.; Hervan, E.M.; Ganjeali, A. Comparative analysis of LEA and Dehydrin genes in response to drought stress in chickpea phonological different. Int. J. Biosci. 2014, 4, 49–57. [Google Scholar]
  30. Krishnamurthy, L.; Kashiwagi, J.; Gaur, P.M.; Upadhyaya, H.D.; Vadez, V. Sources of tolerance to terminal drought in the chickpea (Cicer arietinum L.) minicore germplasm. Field Crops Res. 2010, 119, 322–330. [Google Scholar] [CrossRef] [Green Version]
  31. Upadhyaya, H.D.; Kashiwagi, J.; Varshney, R.K.; Gaur, P.M.; Saxena, K.B.; Krishnamurthy, L.; Gowda, C.L.L.; Pundir, R.P.S.; Chaturvedi, S.K.; Basu, P.S.; et al. Phenotyping chickpeas and pigeonpeas for adaptation to drought. Front. Physiol. 2012, 3, 179. [Google Scholar] [CrossRef] [Green Version]
  32. Awasthi, R.; Gaur, P.; Turner, N.C.; Vadez, V.; Siddique, K.H.M.; Nayyar, H. Effects of individual and combined heat and drought stress during seed filling on the oxidative metabolism and yield of chickpea (Cicer arietinum) genotypes differing in heat and drought tolerance. Crop Pasture Sci. 2017, 68, 823–841. [Google Scholar] [CrossRef]
  33. Devasirvatham, V.; Tan, D.K.Y. Impact of High Temperature and Drought Stresses on Chickpea Production. Agronomy 2018, 8, 145. [Google Scholar] [CrossRef] [Green Version]
  34. Choudhary, M.L.; Tripathi, M.K.; Gupta, N.; Tiwari, S.; Tripathi, N.; Parihar, P.; Pandya, R.K. Screening of pearl millet [Pennisetum glaucum [L] R Br] germplasm lines against drought tolerance based on biochemical traits. Curr. J. Appl. Sci. Technol. 2021, 40, 1–12. [Google Scholar] [CrossRef]
  35. Choudhary, M.L.; Tripathi, M.K.; Tiwari, S.; Pandya, R.K.; Gupta, N.; Tripathi, N.; Parihar, P. Screening of pearl millet [Pennisetum glaucum (L) R Br] germplasm lines for drought tolerance based on morpho-physiological traits and SSR markers. Curr. J. Appl. Sci. Technol. 2021, 40, 46–63. [Google Scholar] [CrossRef]
  36. Raddi, S.; Giannetti, F.; Martini, S.; Farinella, F.; Chirici, G.; Tani, A.; Maltoni, A.; Mariotti, B. Monitoring drought response and chlorophyll content in Quercus by consumer-grade, near-infrared (NIR) camera: A comparison with reflectance spectroscopy. New For. 2022, 53, 241–265. [Google Scholar] [CrossRef]
  37. Sapes, A.; Sala, G. Relative water content consistently predicts drought mortality risk in seedling populations with different morphology, physiology and times to death. Plant Cell Environ. 2021, 44, 3322–3335. [Google Scholar] [CrossRef]
  38. Maqbool, M.A.; Aslam, M.; Ali, H.; Shah, T.M.; Farid, B.; Zaman, Q.U. Drought tolerance indices-based evaluation of chickpea advanced lines under different water treatments. Res. Crops 2015, 16, 336–344. [Google Scholar] [CrossRef]
  39. Kashiwagi, J.; Krishnamurthy, L.; Upadhyaya, H.D.; Krishna, H.; Chandra, S.; Vadez, V.; Serraj, R. Genetic variability of drought-avoidance root traits in the mini-core germplasm collection of chickpeas (Cicer arietinum L). Euphytica 2005, 146, 213–222. [Google Scholar] [CrossRef] [Green Version]
  40. Kashiwagi, J.; Krishnamurthy, L.; Upadhyaya, H.D.; Gaur, P.M. Rapid screening technique for canopy temperature status and its relevance to drought tolerance improvement in chickpea. J. SAT Agric. Res. 2008, 6, 1–4. [Google Scholar]
  41. Krishnamurthy, L.; Kashiwagi, J.; Upadhyaya, H.D.; Serraj, R. Genetic diversity of drought-avoidance root traits in the mini-core germplasm collection of chickpeas. Int. Chick. Pigeonpea Newslett. 2003, 10, 21–24. [Google Scholar]
  42. Ramamoorthy, P.; Lakshmanan, K.; Upadhyaya, H.D.; Vadez, V.; Varshney, R.K. Root traits confer grain yield advantages under terminal drought in chickpea (Cicer arietinum L.). Field Crops Res. 2017, 201, 146–161. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Kumar, N.; Soren, K.R.; Bharadwaj, C.; Sneha Priya, P.R.; Shrivastava, A.K.; Pal, M.; Roorkiwal, M.; Kumar, K.; Patil, B.S.; Soni, A.; et al. Genome-wide transcriptome analysis and physiological variation modulates gene regulatory networks acclimating salinity tolerance in chickpea. Environ. Exp. Bot. 2021, 187, 104478. [Google Scholar] [CrossRef]
  44. Hosseinzadeh, S.R.; Amiri, H.; Ismaili, A. Evaluation of photosynthesis, physiological, and biochemical responses of chickpea (Cicer arietinum L. cv. Pirouz) under water deficit stress and use of vermicompost fertilizer. J. Integrat. Agric. 2018. [Google Scholar] [CrossRef] [Green Version]
  45. Hussain, T.; Akram, Z.; Shabbir, G.; Manaf, A.; Ahmed, M. Identification of drought tolerant Chickpea genotypes through multi trait stability index. Saudi J. Biol. Sci. 2021, 28, 6818–6828. [Google Scholar] [CrossRef]
  46. Salahvarzi, M.; Nasr Esfahani, M.; Shirzadi, N.; Burritt, D.J.; Tran, L.P. Genotype- and tissue-specific physiological and biochemical changes of two chickpea (Cicer arietinum) varieties following a rapid dehydration. Physiol. Plant 2021, 172, 1822–1834. [Google Scholar] [CrossRef]
  47. Thudi, M.; Upadhyaya, H.D.; Rathore, A.; Gaur, P.M.; Krishnamurthy, L.; Roorkiwal, M.; Nayak, S.N.; Chaturvedi, S.K.; Basu, P.S.; Gangarao, N.V.P.R.; et al. Genetic dissection of drought and heat tolerance in chickpea through genomewide and candidate gene-based association mapping approaches. PLoS ONE 2014, 9, e96758. [Google Scholar] [CrossRef] [Green Version]
  48. Palit, P.; Ghosh, R.; Tolani, P.; Tarafdar, A.; Chitikineni, A.; Bajaj, P.; Sharma, M.; Kudapa, H.; Varshney, R.K. Molecular and Physiological Alterations in Chickpea under Elevated CO2 Concentrations. Plant Cell Physiol. 2020, 61, 1449–1463. [Google Scholar] [CrossRef] [PubMed]
  49. Kanca, O.; Bellen, H.J.; Schnorrer, F. Gene Tagging Strategies to Assess Protein Expression, Localization, and Function in Drosophila. Genetics 2017, 207, 389–412, Erratum in Genetics 2017, 207, 1711. [Google Scholar] [CrossRef]
  50. Santiago, C.R.d.N.; Assis, R.d.A.B.; Moreira, L.M.; Digiampietri, L.A. Gene Tags Assessment by Comparative Genomics (GTACG): A User-Friendly Framework for Bacterial Comparative Genomics. Front. Genet. 2019, 10, 725. [Google Scholar] [CrossRef] [Green Version]
  51. Wanga, M.A.; Shimelis, H.; Mashilo, J.; Laing, M.D. Opportunities and challenges of speed breeding: A review. Plant Breed. 2021, 140, 185–194. [Google Scholar] [CrossRef]
  52. Ali, Z.; Merrium, S.; Habib-ur-Rahman, M.; Hakeem, S.; Saddique, M.A.B.; Sher, M.A. Wetting mechanism and morphological adaptation; leaf rolling enhancing atmospheric water acquisition in wheat crop—A review. Environ. Sci. Pollut. Res. 2022, 29, 30967–30985. [Google Scholar] [CrossRef] [PubMed]
  53. Ray, S.; Satya, P. Next generation sequencing technologies for next generation plant breeding. Front. Plant Sci. 2014, 5, 367. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Verma, S.; Gupta, S.; Bandhiwal, N.; Kumar, T.; Bharadwaj, C.; Bhatia, S. High-density linkage map construction and mapping of seed trait QTLs in chickpea (Cicer arietinum L.) using Genotyping-by-Sequencing (GBS). Sci. Rep. 2015, 5, 17512. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Sahu, P.K.; Sao, R.; Mondal, S.; Vishwakarma, G.; Gupta, S.K.; Kumar, V.; Singh, S.; Sharma, D.; Das, B.K. Next Generation Sequencing Based Forward Genetic Approaches for Identification and Mapping of Causal Mutations in Crop Plants: A Comprehensive Review. Plants 2020, 9, 1355. [Google Scholar] [CrossRef]
  56. Wang, Z.; Gerstein, M.; Snyder, M. RNA-Seq: A revolutionary tool for transcriptomics. Nat. Rev. Genet. 2009, 10, 57–63. [Google Scholar] [CrossRef] [PubMed]
  57. Kaur, P.; Gaikwad, K. From Genomes to GENE-omes: Exome Sequencing Concept and Applications in Crop Improvement. Front. Plant Sci. 2017, 8, 2164. [Google Scholar] [CrossRef] [Green Version]
  58. Leo, V.C.; Morgan, N.V.; Bern, D.; Jones, M.L.; Lowe, G.C.; Lordkipanidzé, M.; Drake, S.; Simpson, M.A.; Gissen, P.; Mumford, A.; et al. Use of next-generation sequencing and candidate gene analysis to identify underlying defects in patients with inherited platelet function disorders. J. Thromb. Haemost. 2015, 13, 643–650. [Google Scholar] [CrossRef] [PubMed]
  59. Kulski, J.K.; Suzuki, S.; Ozaki, Y.; Mitsunaga, S.; Inoko, H.; Shiina, T. Phase HLA genotyping by next generation sequencing—A comparison between two massively parallel sequencing bench-top systems, the Roche GS Junior and Ion Torrent PGM. In HLA and Associated Important Diseases; Xi, Y., Ed.; Intech.: Rijeka, Croatia, 2014; pp. 141–181. [Google Scholar]
  60. Rao, M.S.; Van Vleet, T.R.; Ciurlionis, R.; Buck, W.R.; Mittelstadt, S.W.; Blomme, E.A.G.; Liguori, M.J. Comparison of RNA-Seq and Microarray Gene Expression Platforms for the Toxicogenomic Evaluation of Liver from Short-Term Rat Toxicity Studies. Front. Genet. 2019, 9, 636. [Google Scholar] [CrossRef] [PubMed]
  61. Lowe, R.; Shirley, N.; Bleackley, M.; Dolan, S.; Shafee, T. Transcriptomics technologies. PLoS Comput. Biol. 2017, 13, e1005457. [Google Scholar] [CrossRef] [Green Version]
  62. Wang, B.; Kumar, V.; Olson, A.; Ware, D. Reviving the Transcriptome Studies: An Insight into the Emergence of Single-Molecule Transcriptome Sequencing. Front. Genet. 2019, 10, 384. [Google Scholar] [CrossRef] [Green Version]
  63. Bhaskarla, V.; Zinta, G.; Ford, R.; Jain, M.; Varshney, R.K.; Mantri, N. Comparative Root Transcriptomics Provide Insights into Drought Adaptation Strategies in Chickpea (Cicer arietinum L.). Int. J. Mol. Sci. 2020, 21, 1781. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Upasani, M.L.; Limaye, B.M.; Gurjar, G.S.; Kasibhatla, S.M.; Joshi, R.R.; Kadoo, N.Y.; Gupta, V.S. Chickpea-Fusarium oxysporum interaction transcriptome reveals differential modulation of plant defense strategies. Sci. Rep. 2017, 7, 7746. [Google Scholar] [CrossRef] [PubMed]
  65. Aslam, B.; Basit, M.; Nisar, M.A.; Khurshid, M.; Rasool, M.H. Proteomics: Technologies and Their Applications. J. Chromatogr. Sci. 2017, 55, 182–196. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Ahrens, C.H.; Brunner, E.; Qeli, E.; Basler, K.; Aebersold, R. Generating and navigating proteome maps using mass spectrometry. Nat. Rev. Mol. Cell Biol. 2010, 11, 789–801. [Google Scholar] [CrossRef]
  67. Yang, X.; Lu, M.; Wang, Y.; Wang, Y.; Liu, Z.; Chen, S. Response Mechanism of Plants to Drought Stress. Horticulturae 2021, 7, 50. [Google Scholar] [CrossRef]
  68. Mishra, N.; Tripathi, M.K.; Tiwari, S.; Tripathi, N.; Gupta, N.; Sharma, A. Morphological and physiological performance of Indian soybean [Glycine max (L.) Merrill] genotypes in respect to drought. Legume Res. Int. J. 2021. [Google Scholar] [CrossRef]
  69. Mishra, N.; Tripathi, M.K.; Tripathi, N.; Tiwari, S.; Gupta, N.; Sharma, A.; Shrivastav, M.K. Changes in biochemical and antioxidant enzymes activities play significant role in drought tolerance in soybean. Int. J. Agric. Technol. 2021, 17, 1425–1446. [Google Scholar]
  70. Atta, K.; Singh, A.P.; Adhikary, S.; Mondal, S.; Dewanjee, S. Drought Stress: Manifestation and Mechanisms of Alleviation in Plants. In Drought [Working Title] [Internet]; Eyvaz, A.P.M., Albahnasawi, A., Tekbaş, M.M., Gürbulak, M.E., Eds.; IntechOpen: London, UK, 2022; Available online: https://www.intechopen.com/online-first/80834 (accessed on 30 April 2022).
  71. Mishra, N.; Tripathi, M.K.; Tiwari, S.; Tripathi, N.; Ahuja, A.; Sapre, S.; Tiwari, S. Cell suspension culture and in vitro screening for drought tolerance in soybean using poly-ethylene glycol. Plants 2021, 10, 517. [Google Scholar] [CrossRef]
  72. Tiwari, S.; Lata, C.; Chauhan, P.S.; Nautiyal, C.S. Pseudomonas putida attunes morphophysiological, biochemical and molecular responses in Cicer arietinum L. during drought stress and recovery. Plant Physiol. Biochem. 2016, 99, 108–117. [Google Scholar] [CrossRef]
  73. Toker, C.; Canci, H. Selection for drought and heat resistance in chickpea under terminal drought conditions. In Food Legumes for Nutritional Security and Sustainable Agriculture: 4th International Food Legumes Research Conference; Kharkwal, M.C., Ed.; Indian Agricultural Research Institute: New Delhi, India, 2006; pp. 18–22. [Google Scholar]
  74. Varshney, R.K.; Thudi, M.; Pandey, M.K.; Tardieu, F.; Ojiewo, C.; Vadez, V.; Whitbread, A.M.; Siddique, K.H.M.; Nguyen, H.T.; Carberry, P.S.; et al. Accelerating genetic gains in legumes for the development of prosperous smallholder agriculture: Integrating genomics, phenotyping, systems modelling and agronomy. J. Exp. Bot. 2018, 69, 3293–3312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Basu, S.; Ramegowda, V.; Kumar, A.; Pereira, A. Plant adaptation to drought stress. F1000Res 2016, 5, F1000 Faculty Rev-1554. [Google Scholar] [CrossRef]
  76. Comas, L.H.; Becker, S.R.; Cruz, V.M.V.; Byrne, P.F.; Dierig, D.A. Root traits contributing to plant productivity under drought. Front. Plant Sci. 2013, 4, 442. [Google Scholar] [CrossRef] [Green Version]
  77. Martignago, D.; Rico-Medina, A.; Blasco-Escámez, D.; Fontanet-Manzaneque, J.B.; Caño-Delgado, A.I. Drought Resistance by Engineering Plant Tissue-Specific Responses. Front. Plant Sci. 2020, 10, 1676. [Google Scholar] [CrossRef] [PubMed]
  78. Xu, Z.; Zhou, G.; Shimizu, H. Plant responses to drought and rewatering. Plant Signal. Behav. 2010, 5, 649–654. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Bhargava, S.; Sawant, K. Drought stress adaptation: Metabolic adjustment and regulation of gene expression. Plant Breed. 2012, 132, 21–32. [Google Scholar] [CrossRef]
  80. Keskin, M.E.; Terzi, Ö.; Taylan, E.D.; Küçükyaman, D. Meteorological drought analysis using artificial neural networks. Acad. J. 2011, 6, 4469–4477. [Google Scholar]
  81. Berger, J.; Palta, J.; Vadez, V. An integrated framework for crop adaptation to dry environments: Responses to transient and terminal drought. Plant Sci. 2016, 253, 58–67. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Maliro, M.F.A.; MacNeil, D.; Redden, B.; Kollmorgen, J.F.; Pittock, C. Sampling strategies and screening of chickpea (Cicerarietinum L.) germplasm for salt tolerance. Genet. Resour. Crop. Evol. 2008, 55, 53–63. [Google Scholar] [CrossRef]
  83. Kashiwagi, J.; Krishnamurthy, L.; Purushothaman, R.; Upadhyaya, H.D.; Gaur, P.M.; Gowda, C.L.L.; Ito, O.; Varshney, R.K. Scope for improvement of yield under drought through the root traits in chickpea (Cicer arietinum L.). Field Crops Res. 2015, 170, 47–54. [Google Scholar] [CrossRef]
  84. Serraj, R.; Sinclair, T.R. Osmolyte Accumulation: Can It Really Help Increase Crop Yield under Drought Conditions? Plant Cell Environ. 2002, 25, 333–341. [Google Scholar] [CrossRef] [PubMed]
  85. Goulet, B.E.; Roda, F.; Hopkins, R. Hybridization in Plants: Old Ideas, New Techniques. Plant Physiol. 2017, 173, 65–78. [Google Scholar] [CrossRef] [Green Version]
  86. Purushothaman, R.; Krishnamurthy, L.; Upadhyaya, H.D.; Vadez, V.; Varshney, R.K. Shoot traits and their relevance in terminal drought tolerance of chickpea (Cicer arietinum L.). Field Crop. Res. 2016, 197, 10–27. [Google Scholar] [CrossRef] [Green Version]
  87. Leport, L.; Turner, N.C.; Davies, S.L.; Siddique, K.H.M. Variation in pod production and abortion among chickpea cultivars under terminal drought. Eur. J. Agron. 2006, 24, 236–246. [Google Scholar] [CrossRef]
  88. Nayyar, H.; Bains, T.; Kumar, S. Low temperature induced floral abortion in chickpea: Relationship to abscisic acid and cryoprotectants in reproductive organs. Environ. Exp. Bot. 2005, 53, 39–47. [Google Scholar] [CrossRef]
  89. Rahbarian, R.; Nejad, R.K.; Ganjeali, A.; Bagheri, A.; Najafi, F. Drought stress effects on photosynthesis, chlorophyll fluorescence and water relations in tolerant and susceptible chickpea (Cicer arietınum L.) genotypes. Acta Biol. Crac. Ser. Bot. 2011, 53, 47–56. [Google Scholar] [CrossRef]
  90. Gill, S.S.; Tuteja, N. Reactive oxygen species and antioxidant machinery in abiotic stress tolerance in crop plants. Plant Physiol. Biochem. 2010, 48, 909–930. [Google Scholar] [CrossRef] [PubMed]
  91. Kashiwagi, J.; Krishnamurty, L.; Gaur, P.M.; Chandra, S.; Upadhyaya, H.D. Estimation of gene effects of the drought avoidance root characteristicsin chickpea (C. arietinum L.). Field Crops Res. 2008, 105, 64–69. [Google Scholar] [CrossRef] [Green Version]
  92. Purushothaman, R.; Thudi, M.; Krishnamurthy, L.; Upadhyaya, H.D.; Kashiwagi, J.; Gowda, C.L.L.; Varshney, R.K. Association of mid-reproductive stage canopy temperature depression with the molecular markers and grain yields of chickpea (Cicer arietinum L.) germplasm under terminal drought. Field Crops Res. 2015, 174, 1–11. [Google Scholar] [CrossRef] [Green Version]
  93. Sofi, P.A.; Ara, A.; Gull, M.; Rehman, K. Canopy Temperature Depression as an Effective Physiological Trait for Drought Screening. In Drought-Detection and Solutions; IntechOpen: London, UK, 2019. [Google Scholar] [CrossRef] [Green Version]
  94. Karimzadeh, S.H.; Nezami, A.; Nabati, J.; Oskoueian, E.; Ahmadi-Lahijani, M.J. The physiological, biochemical, and molecular modifications of chickpea (Cicer arietinum L.) seedlings under freezing stress. J. Plant Growth Regul. 2021, 41, 1109–1124. [Google Scholar] [CrossRef]
  95. Sivasakthi, K.; Tharanya, M.; Kholová, J.; Wangari Muriuki, R.; Thirunalasundari, T.; Vadez, V. Chickpea Genotypes Contrasting for Vigor and Canopy Conductance Also Differ in Their Dependence on Different Water Transport Pathways. Front. Plant Sci. 2017, 8, 1663. [Google Scholar] [CrossRef] [Green Version]
  96. Krishnamurthy, L.; Kashiwagi, J.; Tobita, S.; Ito, O.; Upadhyaya, H.D.; Gowda, C.L.L.; Gaur, P.M.; Sheshshayee, M.S.; Singh, S.; Vadez, V.; et al. Variation in carbon isotope discrimination and its relationship with harvest index in the reference collection of chickpea germplasm. Funct. Plant Biol. 2013, 14, 1350–1361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Ma, Q.; Behboudian, M.H.; Turner, N.C.; Palta, J.A. Gas exchange by pods and subtending leaves and internal recycling of internal CO2 by pods of chickpea (Cicer arietinum L.) subjected to water stess. J. Exp. Bot. 2001, 52, 123–131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Macar, T.K.; Ekmekci, Y. Alterations in Photochemical and Physiological Activities of Chickpea (Cicer arietinum L.) Cultivars under Drought Stress. J. Agron. Crop. Sci. 2009, 195, 335–346. [Google Scholar] [CrossRef]
  99. Shayla, B.; Inderjit, S.; Satinder, S.; Ashutosh, K.S.G.B.; Sonia, S.; Karan, K.; Kaur, G.S.C.B.; Harsh, N.; Sarvjeet, S. Use of morpho-physiological and biochemical traits to identify sources of drought and heat tolerance in chickpea (Cicer arietinum). Crop. Pasture Sci. 2021, 72, 801–814. [Google Scholar]
  100. Talebi, R.; Ensafi, M.H.; Baghebani, N.; Karami, E.; Mohammadi, K. Physiological responses of chickpea (Cicer arietinum) genotypes to drought stress. Environ. Exp. Biol. 2013, 11, 9–15. [Google Scholar]
  101. Jangpromma, N.; Songsri, P.; Thammasirirak, S.; Jaisil, P. Rapid assessment of chlorophyll content in sugarcane using a SPAD chlorophyll meter across different water stress conditions. Asian J. Plant Sci. 2010, 9, 368–374. [Google Scholar] [CrossRef] [Green Version]
  102. Lamaoui, M.; Jemo, M.; Datla, R.; Bekkaoui, F. Heat and Drought Stresses in Crops and Approaches for Their Mitigation. Front. Chem. 2018, 6, 26. [Google Scholar] [CrossRef] [PubMed]
  103. Devasirvatham, V.; Tan, D.K.Y.; Gaur, P.M.; Raju, T.N.; Trethowan, R.M. High temperature tolerance in chickpea and its implications for plant improvement. Crop. Pasture Sci. 2012, 63, 419–428. [Google Scholar] [CrossRef] [Green Version]
  104. Chen, Y.; Ghanem, M.E.; Siddique, K.H. Characterising root trait variability in chickpea (Cicer arietinum L.) germplasm. J. Exp. Bot. 2017, 68, 1987–1999. [Google Scholar] [CrossRef] [Green Version]
  105. Badhan, S.; Kole, P.; Ball, A.; Mantri, N. RNA sequencing of leaf tissues from two contrasting chickpea genotypes reveals mechanisms for drought tolerance. Plant Physiol. Biochem. 2018, 129, 295–304. [Google Scholar] [CrossRef]
  106. Ahmar, S.; Gill, R.A.; Jung, K.H.; Faheem, A.; Qasim, M.U.; Mubeen, M.; Zhou, W. Conventional and Molecular Techniques from Simple Breeding to Speed Breeding in Crop Plants: Recent Advances and Future Outlook. Int. J. Mol. Sci. 2020, 21, 2590. [Google Scholar] [CrossRef]
  107. Stefaniak, T.; McPhee, K. Comparison of Hybridization Techniques in Chickpea. Crop. Sci. 2017, 57, 843–846. [Google Scholar] [CrossRef]
  108. Deokar, A.A.; Tar’an, B. Genome-wide analysis of the aquaporin gene family in chickpea (Cicer arietinum L.). Front. Plant Sci. 2016, 7, 1802. [Google Scholar] [CrossRef] [Green Version]
  109. Deokar, A.A.; Kondawar, V.; Kohli, D.; Aslam, M.; Jain, P.K.; Karuppayil, S.M.; Varshney, R.K.; Srinivasan, R. The CarERF genes in chickpea (Cicerarietinum L.) and the identification of CarERF116 as abiotic stress responsive transcription factor. Funct. Integr. Genom. 2015, 15, 27–46. [Google Scholar] [CrossRef] [PubMed]
  110. Hamwieh, A.; Imtiaz, M.; Malhotra, R.S. Multi-environment QTL analyses for drought-related traits in a recombinant inbred population of chickpea (Cicer arietinum L.). Theor. Appl. Genet. 2013, 126, 1025–1038. [Google Scholar] [CrossRef] [PubMed]
  111. Singh, V.K.; Khan, A.W.; Jaganathan, D.; Thudi, M.; Roorkiwal, M.; Takagi, H.; Garg, V.; Kumar, V.; Chitikineni, A.; Gaur, P.M.; et al. QTL-seq for rapid identification of candidate genes for 100-seed weight and root/total plant dry weight ratio under rainfed conditions in chickpea. Plant. Biotechnol. J. 2016, 14, 2110–2119. [Google Scholar] [CrossRef] [Green Version]
  112. Kudapa, H.; Garg, V.; Chitikineni, A.; Varshney, R.K. The RNA-Seqbased high resolution gene expression atlas of chickpea (Cicer arietinum L.) reveals dynamic spatiotemporal changes associated with growth and development. Plant Cell Environ. 2018, 41, 2209–2225. [Google Scholar] [CrossRef] [Green Version]
  113. Kumar, M.; Chauhan, A.S.; Yusuf, M.A.; Sanyal, I.; Chauhan, P.S. Transcriptome sequencing of chickpea (Cicer arietinum L.) genotypes for identifcation of drought-responsive genes under drought stress condition. Plant Mol. Biol. Rep. 2019, 37, 186–203. [Google Scholar] [CrossRef]
  114. Gupta, S.; Singh, A.; Singh, P.; Kewat, R.N. Effect of drought stress or carbohydrate content in drought tolerant and susceptible chickpea genotypes. J. Crop. Sci. Biotechnol. 2015, 4, 35–38. [Google Scholar]
  115. Sachdeva, S.; Bharadwaj, C.; Singh, R.K.; Jain, P.K.; Patil, B.S.; Roorkiwal, M.; Varshney, R. Characterization of ASR gene and its role in drought tolerance in chickpea (Cicer arietinum L.). PLoS ONE 2020, 15, e0234550. [Google Scholar] [CrossRef]
  116. Kooyers, N.J. The evolution of drought escape and avoidance in natural herbaceous populations. Plant Sci. 2015, 234, 155–162. [Google Scholar] [CrossRef] [PubMed]
  117. Siddique, K.H.M.; Loss, S.P.; Thomson, B.D. Cool season grain legumes in dryland Mediterranean environments of Western Australia: Significance of early flowering. In Management of Agricultural Drought; Saxena, N.P., Ed.; Science Publishers: Enfield, NH, USA, 2003; pp. 151–161. [Google Scholar]
  118. Saeed, M.; Francis, C.A. Yield stability in relation to maturity in grain sorghum. Crop. Sci. 1983, 23, 683–687. [Google Scholar] [CrossRef]
  119. Richards, M.F.; Maphosa, L.; Preston, A.L. Impact of Sowing Time on Chickpea (Cicer arietinum L.) Biomass Accumulation and Yield. Agronomy 2022, 12, 160. [Google Scholar] [CrossRef]
  120. Shimray, P.U.; Bajaj, D.; Shrivastava, R.; Daware, A.; Upadhyaya, H.D.; Kumar, R.; Bhardwaj, C.; Tyagi, A.K.; Parida, S.K. Identifying Transcription Factor Genes Associated with Yield Traits in Chickpea. Plant Mol. Biol. Rep. 2017, 35, 562–574. [Google Scholar] [CrossRef]
  121. Sabaghpour, S.H.; Kumar, J.; Rao, T.N. Inheritance of growth vigor and its association with other characters in chickpea. Plant Breed. 2003, 122, 542–544. [Google Scholar] [CrossRef] [Green Version]
  122. Sun, F.; Chen, Q.; Chen, Q.; Jiang, M.; Gao, W.; Qu, Y. Screening of Key Drought Tolerance Indices for Cotton at the Flowering and Boll Setting Stage Using the Dimension Reduction Method. Front. Plant Sci. 2021, 12, 619926. [Google Scholar] [CrossRef]
  123. Sabaghpour, S.H.; Mahmodi, A.A.; Saeed, A.; Kamel, M.; Malhotra, R.S. Study on chickpea drought tolerance lines under dryland condition of Iran. Indian J. Crop. Sci. 2006, 1, 70–73. [Google Scholar]
  124. Dumanović, J.; Nepovimova, E.; Natić, M.; Kuča, K.; Jaćević, V. The Significance of Reactive Oxygen Species and Antioxidant Defense System in Plants: A Concise Overview. Front. Plant Sci. 2021, 11, 552969. [Google Scholar] [CrossRef]
  125. Levitt, J. Responses of Plants to Environmental Stresses. Volume II. Water, Radiation, Salt, and Other Stresses; Academic Press: Cambridge, MA, USA, 1980. [Google Scholar]
  126. Kamanga, R.M.; Mbega, E.; Ndakidemi, P. Drought Tolerance Mechanisms in Plants: Physiological Responses Associated with Water Deficit Stress in Solanum lycopersicum. Adv. Crop. Sci. Technol. 2018, 6, 362. [Google Scholar] [CrossRef]
  127. Pastori, G.; Foyer, C.H.; Mullineaux, P. Low temperature-induced changes in the distribution of H2O2 and antioxidants between the bundle sheath and mesophyll cells of maize leaves. J. Exp. Bot. 2000, 51, 107–113. [Google Scholar] [CrossRef] [Green Version]
  128. Li, H.; Ma, X.; Lu, Y.; Ren, R.; Cui, B.; Si, B. Growing deep roots has opposing impacts on the transpiration of apple trees planted in subhumid loess region. Agric. Water Manag. 2021, 258, 107207. [Google Scholar] [CrossRef]
  129. Maqbool, M.A.; Aslam, M.; Ali, H.; Shah, T.M.; Atta, B.M. GGE biplot analysis-based selection of superior chickpea (Cicerarietinum L.) inbred lines under variable water environments. Pak. J. Bot. 2015, 47, 1901–1908. [Google Scholar]
  130. Sohrabi, Y.; Heidari, G.; Weisany, W.; Ghasemi Golezani, K.; Mohammadi, K. Some physiological responses of chickpea cultivars to arbuscular mycorrhiza under drought stress. Russ. J. Plant Physiol. 2012, 59, 708–716. [Google Scholar] [CrossRef]
  131. Kashiwagi, J.; Krishnamurthy, L.; Crouch, J.H.; Serraj, R. Variability of root length density and its contributions to seed yield in chickpea (Cicer arietinum L) under terminal drought stress. Field Crops Res. 2006, 95, 171–181. [Google Scholar] [CrossRef]
  132. Gaur, P.M.; Krishnamurthy, L.; Kashiwagi, J. Improving drought avoidance root traits in chickpea (Cicer arietinum L.) current status of research at ICRISAT. Plant Prod. Sci. 2008, 11, 3–11. [Google Scholar] [CrossRef] [Green Version]
  133. Davies, W.; Zhang, J. Root signals and the regulation of growth and development of plants in dry soil. Ann. Rev. Plant Physiol. Plant Mol. Biol. 1991, 42, 55–76. [Google Scholar] [CrossRef]
  134. Pierik, R.; Sasidharan, R.; Voesenek, L.A.C.J. Growth control by ethylene: Adjusting phenotypes to the environment. J. Plant Growth Regul. 2007, 26, 188–200. [Google Scholar] [CrossRef] [Green Version]
  135. Zaman-Allah, M.; Jenkinson, D.M.; Vandez, V. Chickpea genotypes contrasting for seed yield under terminal drought stress in the field differ for traits related to the control of water use. Funct. Plant Biol. 2011, 38, 270–281. [Google Scholar] [CrossRef] [Green Version]
  136. Pang, J.; Turner, N.C.; Du, Y.-L.; Colmer, T.D.; Siddique, K.H.M. Pattern of Water Use and Seed Yield under Terminal Drought in Chickpea Genotypes. Front. Plant Sci. 2017, 8, 1375. [Google Scholar] [CrossRef] [Green Version]
  137. Zaman-Allah, M.; Jenkinson, D.M.; Vandez, V. A conservative pattern of water use, rather than deep or profuse rooting, is critical for terminal drought tolerance of chickpea. J. Exp. Bot. 2011, 38, 270–281. [Google Scholar] [CrossRef] [Green Version]
  138. Kashiwagi, J.; Krishnamurthy, L.; Singh Sube Gaur, P.M.; Upadhyaya, H.D.; Panwar, J.D.S.; Basu, P.S.; Ito, O.; Tobita, S. Relationship between transpiration efficiency and carbon isotope discrimination in chickpea (Cicer arietinum L.). J. SAT Agric. Res. 2006, 2, 1–3. [Google Scholar]
  139. Nadal-Sala, D.; Grote, R.; Birami, B.; Knüver, T.; Rehschuh, R.; Schwarz, S.; Ruehr, N.K. Leaf Shedding and Non-Stomatal Limitations of Photosynthesis Mitigate Hydraulic Conductance Losses in Scots Pine Saplings During Severe Drought Stress. Front. Plant Sci. 2021, 12, 715127. [Google Scholar] [CrossRef] [PubMed]
  140. Li, Y.; Li, H.; Li Zhang, S. Improving water-use efficiency by decreasing stomatal conductance and transpiration rate to maintain higher ear photosynthetic rate in drought-resistant wheat. Crop. J. 2017, 5, 231–239. [Google Scholar] [CrossRef]
  141. Flexas, J.; Medrano, H. Drought-inhibition of photosynthesis in C3 plants: Stomatal and non-stomatal limitations revisited. Ann. Bot. 2002, 89, 183–189. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Cornic, G. Drought stress inhibits photosynthesis by decreasing stomatal aperture not by affecting ATP synthesis. Trends Plant Sci. 2000, 5, 187–188. [Google Scholar] [CrossRef]
  143. Flexas, J.; Galmes, J.; Ribas-Carbo, M.; Medrano, H. The effects of water stress on plant respiration. In Plant Respiration: From Cell to Ecosystem; Lambers, H., Ribas-Carbo, M., Eds.; Advances in Photosynthesis and Respiration Series; Springer: Dordrecht, The Netherlands, 2005; Chapter 6; Volume 18, pp. 85–94. [Google Scholar]
  144. Farooq, M.; Gogoi, N.; Barthakur, S.; Baroowa, B.; Bharadwaj, N.; Alghamdi, S.S.; Siddique, K.H.M. Drought stress in grain legumes during reproduction and grain filling. J. Agron. Crop. Sci. 2017, 203, 81–102. [Google Scholar] [CrossRef]
  145. Xiao, Q.; Bai, X.; Zhang, C.; He, Y. Advanced high-throughput plant phenotyping techniques for genome-wide association studies: A review. J. Adv. Res. 2022, 35, 215–230. [Google Scholar] [CrossRef]
  146. Nayyar, H.; Singh, S.; Kaur, S.; Kumar, S.; Upadhyaya, H.D. Differential sensitivity of macrocarpa and microcarpa types of chickpea (Cicer arietinum L.) to water stress: Association of contrasting stress response with oxidative injury. J. Integr. Plant Biol. 2006, 48, 1318–1329. [Google Scholar] [CrossRef]
  147. Masle, J.; Gilmore, S.R.; Farquhar, G.D. The ERECTA gene regulates plant transpiration efficiency in Arabidopsis. Nature 2005, 436, 866–870. [Google Scholar] [CrossRef]
  148. Saxena, N.P. Management of drought in chickpea-holistic approach. In Management of Agricultural Drought-Agronomic and Genetic Options; Saxena, N.P., Ed.; Oxford and IBH Publishing Co. Pvt. Ltd.: New Delhi, India, 2003; pp. 103–122. [Google Scholar]
  149. Hemati, A.; Mofidi-Chelan, M.; Amirifar, A.; Moghiseh, E.; Asgari Lajayer, B. Drought Tolerance Mechanisms in Crop Plants. In Plant Stress Mitigators; Vaishnav, A., Arya, S., Choudhary, D.K., Eds.; Springer: Singapore, 2022. [Google Scholar] [CrossRef]
  150. Imtiaz, M.; Malhotra, R.S. Reduce stress: Breed for drought tolerance. ICARDA Caravan Sci. Food Sec. 2009, 26, 34–36. [Google Scholar]
  151. Close, T.J. Dehydrins: Emergence of a biochemical role of a family of plant dehydration proteins. Physiol. Plant 1996, 97, 795–803. [Google Scholar] [CrossRef]
  152. Hara, M.; Terashima, S.; Kuboi, T. Characterization and cryoprotective activity of cold-responsive dehydrin from Citrus unshiu. J. Plant Physiol. 2001, 158, 1333–1339. [Google Scholar] [CrossRef]
  153. Hung, S.H.; Yu, C.W. Hydrogen peroxide functions as a stress signal in plants. Bot. Bull. Acad. Sinica 2005, 41, 1–10. [Google Scholar]
  154. Chen, Z.; Gallie, D.R. The ascorbic acid redox state controls guard cell signaling and stomatal movement. Plant Cell. 2004, 16, 1143–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Sharma, P.; Jha, A.B.; Dubey, R.S.; Pessarakli, M. Reactive Oxygen Species, Oxidative Damage, and Antioxidative Defense Mechanism in Plants under Stressful Conditions. J. Bot. 2012, 2012, 217037. [Google Scholar] [CrossRef] [Green Version]
  156. 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]
  157. Huang, H.; Ullah, F.; Zhou, D.-X.; Yi, M.; Zhao, Y. Mechanisms of ROS Regulation of Plant Development and Stress Responses. Front. Plant Sci. 2019, 10, 800. [Google Scholar] [CrossRef]
  158. Gupta, P.K.; Varshney, R.K.; Prasad, M. Molecular markers: Principles and methodology. In Molecular Techniques in Crop Improvement; Jain, S.M., Ahloowalia, B.S., Brar, D.S., Eds.; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2002; pp. 9–54. [Google Scholar]
  159. Pushpavalli, R.; Krishnamurthy, L.; Thudi, M.; Gaur, P.M.; Rao, M.V.; Siddique, K.H.M.; Colmer, T.D.; Turner, N.C.; Varshney, R.K.; Vadez, V. Two key genomic regions harbour QTLs for salinity tolerance in ICCV 2 JG 11 derived chickpea (Cicerarietinum L.) recombinant inbred lines. BMC Plant Biol. 2015, 15, 124. [Google Scholar] [CrossRef] [Green Version]
  160. Li, Y.; Ruperao, P.; Batley, J.; Edwards, D.; Khan, T.; Colmer, T.D.; Pang, J.; Siddique, K.H.M.; Sutton, T. Investigating Drought Tolerance in Chickpea Using Genome-Wide Association Mapping and Genomic Selection Based on Whole-Genome Resequencing Data. Front. Plant Sci. 2018, 9, 190. [Google Scholar] [CrossRef] [Green Version]
  161. Kalra, N.; Chakraborty, D.; Sharma, A.; Rai, H.K.; Jolly, M.; Chander, S.; Kumar, P.R.; Bhadraray Barman, D.; Mittal, R.B.; Lal, M.; et al. Effect of increasing temperature on yield of some winter crops in northwest India. Curr. Sci. 2008, 94, 82–88. [Google Scholar]
  162. Monneveux, P.; Ribaut, J.-M. Secondary traits for drought tolerance improvement in cereals. In Drought Adaptation in Cereals; Ribaut, J.-M., Ed.; The Haworth Press Inc.: Binghamton, NY, USA, 2006; pp. 97–143. [Google Scholar]
  163. Shah, T.M.; Imran, M.; Atta, B.M.; Ashraf, M.Y.; Hameed, A.; Waqar, I.; Shafiq, M.; Hussain, K.; Naveed, M.; Aslam, M.; et al. Selection and screening of drought tolerant high yielding chickpea genotypes based on physio-biochemical indices and multi-environmental yield trials. BMC Plant Biol. 2020, 20, 171. [Google Scholar] [CrossRef] [PubMed]
  164. Gong, H.; Zhu, X.; Chen, K.; Wang, S.; Zhang, C. Silicon alleviates oxidative damage of wheat plants in pots under drought. Plant Sci. 2005, 169, 313–321. [Google Scholar] [CrossRef]
  165. Hasegawa, P.M.; Bressan, R.A.; Zhu, J.K.; Bohnert, H.J. Plant cellular and molecular responses to high salinity. Annu. Rev. Plant. Biol. 2000, 51, 463–499. [Google Scholar] [CrossRef] [Green Version]
  166. Fazeli, F.; Ghorbanli, M.; Niknam, V. Effect of drought on biomass, protein content, lipid peroxidation and antioxidant enzymes in two sesame cultivars. Biol. Plant. 2007, 51, 98–103. [Google Scholar] [CrossRef]
  167. Lima, A.L.S.; DaMatta, F.M.; Pinheiro, H.A.; Totola, M.R.; Loureiro, M.E. Photo- chemical responses and oxidative stress in two clones of Coffea canephora under water deficit conditions. Environ. Exp. Bot. 2002, 47, 239–247. [Google Scholar] [CrossRef]
  168. Chew, O.; Whelan, J.; Miller, A.H. Molecular definition of the ascorbate-gluta- thione cycle in Arabidopsis mitochondria reveals dual targeting of antioxidant defences in plants. J. Biol. Chem. 2003, 278, 46869–46877. [Google Scholar] [CrossRef] [Green Version]
  169. Kaur, D.; Grewal, S.K.; Kaur, J.; Singh, S. Differential proline metabolism in vegetative and reproductive tissues determines drought tolerance in chickpea. Biol. Plant. 2017, 61, 359–366. [Google Scholar] [CrossRef]
  170. Gurrieri, L.; Merico, M.; Trost, P.; Forlani, G.; Sparla, F. Impact of drought on soluble sugars and free proline content in selected Ara bidopsis mutants. Biology 2020, 9, 367. [Google Scholar] [CrossRef]
  171. Iqbal, S.; Wang, X.; Mubeen, I.; Kamran, M.; Kanwal, I.; Díaz, G.A.; Abbas, A.; Parveen, A.; Atiq, M.N.; Alshaya, H.; et al. Phytohormones Trigger Drought Tolerance in Crop Plants: Outlook and Future Perspectives. Front. Plant Sci. 2022, 12, 799318. [Google Scholar] [CrossRef]
  172. Taiz, L.; Zeiger, E. Plant Physiology, 4th ed.; Sinauer Associates Inc. Publishers: Sunderland, MA, USA, 2006. [Google Scholar]
  173. Dubois, M.; Van den Broeck, L.; Inzé, D. The Pivotal Role of Ethylene in Plant Growth. Trends Plant Sci. 2018, 23, 311–323. [Google Scholar] [CrossRef] [Green Version]
  174. Nilsen, E.T.; Orcutte, D.M. Phytohormones and plant responses to stress. In Physiology of Plant under Stress: Abiotic Factors; John Wiley and Sons: New York, NY, USA, 1996; pp. 183–198. [Google Scholar]
  175. Taylor, I.B. Genetics of ABA synthesis. In Abscisic Acid: Physiology and Biochemistry; Davies, W.J., Jones, H.G., Eds.; Bios Scientific Publishers Ltd.: Oxford, UK, 1991; pp. 23–38. [Google Scholar]
  176. Kasukabe, Y.; He, L.; Nada, K.; Misawa, S.; Ihara, I.; Tachibana, S. Overexpression of spermidine synthase enhances tolerance to multiple environmental stresses and up- regulates the expression of various stress-regulated genes in transgenic Arabidopsis thaliana. Plant Cell. Physiol. 2004, 45, 712–722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Salimath, P.M.; Toker, C.; Sandhu, J.S.; Kumar, J.; Suma, B.; Yadav, S.S.; Bahl, P.N. Conventional breeding methods. In Chickpea Breeding and Management; Yadav, S.S., Redden, R.J., Chen, W., Sharma, B., Eds.; CABI: Wallingford, UK, 2007; pp. 369–390. [Google Scholar]
  178. Holme, I.B.; Gregersen, P.L.; Brinch-Pedersen, H. Induced Genetic Variation in Crop Plants by Random or Targeted Mutagenesis: Convergence and Differences. Front. Plant Sci. 2019, 10, 1468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Micke, A. Genetic improvement of food legumes in developing countries by mutation induction. In World Crops: Cool Season Food Legumes; Summerfield, R.J., Ed.; Kluwer Academic: Dordrecht, The Netherlands, 1988; pp. 1031–1047. [Google Scholar]
  180. Amri-Tiliouine, W.; Laouar, M.; Abdelguerfi, A.; Jankowicz-Cieslak, J.; Jankuloski, L.; Till, B.J. Genetic Variability Induced by Gamma Rays and Preliminary Results of Low-Cost TILLING on M2 Generation of Chickpea (Cicer arietinum L.). Front. Plant Sci. 2018, 9, 1568. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Kumar, S.; Katna, G.; Sharma, N. Mutation breeding in chickpea. Adv. Plants Agric. Res. 2019, 9, 355–362. [Google Scholar] [CrossRef]
  182. Dua, R.P.; Chaturvedi, S.K.; Shiv, S. Reference Varieties of chickpea for IPR Regime; IIPR: Kanpur, India, 2001. [Google Scholar]
  183. Ma, P.; Zhang, X.; Chen, L.; Zhao, Q.; Zhang, Q.; Wang, Z.; Tang, H.; Yu, Q.; Zhang, M.; Ming, R.; et al. Comparative analysis of sucrose phosphate synthase (SPS) gene family between Saccharum officinarum and Saccharum spontaneum. BMC Plant Biol. 2020, 20, 422. [Google Scholar] [CrossRef]
  184. Harb, A.; Krishnan, A.; Ambavaram, M.M.; Pereira, A. Molecular and physiological analysis of drought stress in Arabidopsis reveals early responses leading to acclimation in plant growth. Plant Physiol. 2010, 154, 1254–1271. [Google Scholar] [CrossRef] [Green Version]
  185. Yao, J.; Sun, D.; Cen, H.; Xu, H.; Weng, H.; Yuan, F.; He, Y. Phenotyping of Arabidopsis Drought Stress Response Using Kinetic Chlorophyll Fluorescence and Multicolor Fluorescence Imaging. Front. Plant Sci. 2018, 9, 603. [Google Scholar] [CrossRef]
  186. Granier, C.; Aguirrezabal, L.; Chenu, K.; Cookson, S.J.; Dauzat, M.; Hamard, P.; Thioux, J.J.; Rolland, G.; Bouchier-Combaud, S.; Lebaudy, A. PHENOPSIS, an automated platform for reproducible phenotyping of plant responses to soil water deficit in Arabidopsis thaliana permitted the identification of an accession with low sensitivity to soil water deficit. New Phytol. 2006, 169, 623–635. [Google Scholar] [CrossRef]
  187. Bouchabke-Coussa, O.; Quashie, M.L.; Seoane-Redondo, J.; Fortabat, M.N.; Gery, C.; Yu, A.; Linderme, D.; Trouverie, J.; Granier, F.; Téoulé, E.; et al. ESKIMO1 is a key gene involved in water economy as well as cold acclimation and salt tolerance. BMC Plant Biol. 2008, 8, 125. [Google Scholar] [CrossRef] [Green Version]
  188. Bhatnagar-Mathur, P.; Vadez, V.; Devi, M.; Lavanya, M.; Vani, G.; Sharma, K. Genetic engineering of chickpea (Cicer arietinum L.) with the P5CSF129A gene for osmoregulation with implications on drought tolerance. Mol. Breed. 2009, 23, 591–606. [Google Scholar] [CrossRef] [Green Version]
  189. Kumar, T.; Bharadwaj, C.; Rizvi, A.H.; Sarker, A.; Tripathi, S.; Alam, A.; Chauhan, S.K. Chickpea Landraces: A Valuable and Divergent Source for Drought Tolerance. Int. J. Trop. Agric. 2015, 33, 1–6. [Google Scholar]
  190. Amiteye, S. Basic concepts and methodologies of DNA marker systems in plant molecular breeding. Heliyon 2021, 7, e08093. [Google Scholar] [CrossRef]
  191. Tiwari, S.; Tripathi, N.; Tsuji, K.; Tantwai, K. Genetic diversity and population structure of Indian soybean (Glycine max (L.) Merr.) as revealed by microsatellite markers. Physiol. Mol. Biol. Plants 2019, 25, 953–964. [Google Scholar] [CrossRef] [PubMed]
  192. Kachare, S.; Tiwari, S.; Tripathi, N.; Thakur, V.V. Assessment of genetic diversity of soybean (Glycine max (L.) Merr.) genotypes using qualitative traits and microsatellite makers. Agric. Res. 2019, 9, 23–34. [Google Scholar] [CrossRef]
  193. Gupta, P.K.; Rustgi, S.; Mir, R.R. Array-based high-throughput DNA markers for crop improvement. Heredity 2008, 101, 5–18. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Serret, M.D.; Udupa, S.M.; Weigand, F. Assessment of genetic diversity of cultivated chickpea using microsatellite-derived RFLP markers: Implications for origin. Plant Breed. 2006, 116, 573–578. [Google Scholar] [CrossRef]
  195. Iruela, M.; Rubio, J.; Cubero, J.I.; Gil, J.; Mill´an, T. Phylogenetic analysis in the genus Cicer and cultivated chickpea using RAPD and ISSR markers. Theor. Appl. Genet. 2002, 104, 643–651. [Google Scholar] [CrossRef]
  196. Yadav, P.; Tomar, V.S.; Mishra, J.K.; Chauhan, A.K.S. Relationship among annual Cicer species using random amplified polymorphic DNA markers. Indian J. Life Sci. Gale Acad. 2012. Available online: link.gale.com/apps/doc/A357968924/AONE?u=anon~91e4fa6e&sid=googleScholar&xid=cbcb2ca8 (accessed on 20 October 2022).
  197. Tripathi, N.; Chouhan, D.S.; Saini, N.; Tiwari, S. Assessment of genetic variations among highly endangered medicinal plant Bacopa monnieri (L.) from Central India using RAPD and ISSR analysis. 3 Biotech 2012, 2, 327–336. [Google Scholar] [CrossRef]
  198. Singh, R.; Singhal, V.; Randhawa, G.J. Molecular Analysis of Chickpea (Cicer arietinum L) Cultivars Using AFLP and STMS Markers. J. Plant Biochem. Biotechnol. 2008, 17, 167–171. [Google Scholar] [CrossRef]
  199. Hajibarat, Z.; Saidi, A.; Hajibarat, Z.; Talebi, R. Characterization of genetic diversity in chickpea using SSR markers, Start Codon Targeted Polymorphism (SCoT) and Conserved DNA-Derived Polymorphism (CDDP). Physiol. Mol. Biol. Plants 2015, 21, 365–373. [Google Scholar] [CrossRef]
  200. Gajbhiye, P.; Yadav, A.S. Molecular markers in chickpea improvement: A review. Plant Arch. 2012, 12, 589–597. [Google Scholar]
  201. Bharadwaj, C.; Chauhan, S.K.; Rajguru, G.; Srivastava, R.; Satyavathi, C.T.; Yadav, S. Diversity analysis of chickpea (Cicer arietinum L.) using STMS markers. Indian J. Agric. Sci. 2010, 80, 947–951. [Google Scholar]
  202. Varshney, R.K.; Kudapa, H.; Roorkiwal, M.; Thudi, M.; Pandey, M.K.; Saxena, R.K.; Chamarthi, S.K.; Mohan, S.M.; Mallikarjuna, N.; Upadhyaya, H.; et al. Advances in genetics and molecular breeding of three legume crops of semi-arid tropics using next generation sequencing and high-throughput genotyping technologies. J. Biosci. 2012, 37, 811–820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Nayak, S.N.; Zhu, H.; Varghese, N.; Datta, S.; Choi, H.K.; Horres, R.; Jüngling, R.; Singh, J.; Kavi Kishor, P.B.; Sivaramakrishnan, S.; et al. Integration of novel SSR and gene-based SNP marker loci in the chickpea genetic map and establishment of new anchor points with Medicago truncatula genome. Theor. Appl. Genet. 2010, 120, 1415–1441. [Google Scholar] [CrossRef] [Green Version]
  204. Seyedimoradi, H.; Talebi, R.; Kanouni, H.; Naji, A.M.; Karami, E. Genetic diversity and population structure analysis of chickpea (Cicer arietinum L.) advanced breeding lines using whole-genome DArTseq-generated SilicoDArT markers. Braz. J. Bot. 2020, 43, 541–549. [Google Scholar] [CrossRef]
  205. Ahmed, S.M.; Alsamman, A.M.; Jighly, A.; Mubarak, M.H.; Al-Shamaa, K.; Istanbuli, T.; Momtaz, O.A.; Allali, A.E.; Hamwieh, A. Genome-wide association analysis of chickpea germplasms differing for salinity tolerance based on DArTseq markers. PLoS ONE 2021, 16, e0260709. [Google Scholar] [CrossRef]
  206. Hiremath, P.J.; Kumar, A.; Penmetsa, R.V.; Farmer, A.; Schlueter, J.A.; Chamarthi, S.K.; Whaley, A.M.; Carrasquilla-Garcia, N.; Gaur, P.M.; Upadhyaya, H.D.; et al. Large-scale development of cost-effective SNP marker assays for diversity assessment and genetic mapping in chickpea and comparative mapping in legumes. Plant Biotechnol. J. 2012, 10, 716–732. [Google Scholar] [CrossRef] [Green Version]
  207. Varshney, R.K. Exciting journey of 10 years from genomes to felds and markets: Some success stories of genomics-assisted breeding in chickpea, pigeonpea and groundnut. Plant Sci. 2016, 242, 98–107. [Google Scholar] [CrossRef] [Green Version]
  208. Yadav, G.; Jayaswal, D.; Jayaswall, K.; Bhandawat, A.; Singh, A.N.; Tilgam, J.; Rai, A.K.; Chaturvedi, R.; Kumar, A.; Kumar, S.; et al. Identification and characterization of chickpea genotypes for early flowering and higher seed germination through molecular markers. Mol. Biol. Rep. 2022, 49, 6181–6188. [Google Scholar] [CrossRef]
  209. Rajkumar, M.S.; Garg, R.; Jain, M. Genome-wide discovery of DNA polymorphisms via resequencing of chickpea cultivars with contrasting response to drought stress. Physiol. Plant 2022, 174, e13611. [Google Scholar] [CrossRef]
  210. Basu, U.; Srivastava, R.; Bajaj, D.; Thakro, V.; Daware, A.; Malik, N.; Upadhyaya, H.D.; Parida, S.K. Genome-wide generation and genotyping of informative SNPs to scan molecular signatures for seed yield in chickpea. Sci. Rep. 2018, 8, 13240. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Rajkumar, M.S.; Garg, R.; Jain, M. Genome-wide discovery of DNA polymorphisms among chickpea cultivars with contrasting seed size/weight and their functional relevance. Sci. Rep. 2018, 8, 16795. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Manchikatla, P.K.; Kalavikatte, D.; Mallikarjuna, B.P.; Palakurthi, R.; Khan, A.W.; Jha, U.C.; Bajaj, P.; Singam, P.; Chitikineni, A.; Varshney, R.K.; et al. MutMap Approach Enables Rapid Identification of Candidate Genes and Development of Markers Associated with Early Flowering and Enhanced Seed Size in Chickpea (Cicer arietinum L.). Front. Plant Sci. 2021, 12, 688694. [Google Scholar] [CrossRef] [PubMed]
  213. Srivastava, R.; Singh, M.; Bajaj, D.; Parida, S.K. A High-Resolution In Del (Insertion–Deletion) Markers-Anchored Consensus Genetic Map Identifies Major QTLs Governing Pod Number and Seed Yield in Chickpea. Front. Plant Sci. 2016, 7, 1362. [Google Scholar] [CrossRef] [Green Version]
  214. Mallikarjuna, B.P.; Samineni, S.; Thudi, M.; Sajja, S.B.; Khan, A.W.; Patil, A.; Viswanatha, K.P.; Varshney, R.K.; Gaur, P.M. Molecular Mapping of Flowering Time Major Genes and QTLs in Chickpea (Cicer arietinum L.). Front. Plant Sci. 2017, 8, 1140. [Google Scholar] [CrossRef] [Green Version]
  215. Barmukh, R.; Soren, K.R.; Madugula, P.; Gangwar, P.; Shanmugavadivel, P.S.; Bharadwaj, C.; Konda, A.K.; Chaturvedi, S.K.; Bhandari, A.; Rajain, K.; et al. Construction of a high-density genetic map and QTL analysis for yield, yield components and agronomic traits in chickpea (Cicer arietinum L.). PLoS ONE 2021, 16, e0251669. [Google Scholar] [CrossRef]
  216. Rezaei, M.K.; Deokar, A.A.; Arganosa, G.; Roorkiwal, M.; Pandey, S.K.; Warkentin, T.D.; Varshney, R.K.; Tar An, B. Mapping Quantitative Trait Loci for Carotenoid Concentration in Three F2 Populations of Chickpea. Plant Genome 2019, 12, 67. [Google Scholar] [CrossRef] [Green Version]
  217. Varshney, R.K.; Thudi, M.; Nayak, S.N.; Gaur, P.M.; Kashiwagi, J.; Krishnamurthy, L.; Jaganathan, D.; Koppolu, J.; Bohra, A.; Tripathi, S.; et al. Genetic dissection of drought tolerance in chickpea (Cicer arietinum L.). Theor. Appl. Genet. 2014, 127, 445–462. [Google Scholar] [CrossRef]
  218. Varshney, R.K.; Song, C.; Saxena, R.K.; Azam, S.; Yu, S.; Sharpe, A.G.; Cannon, S.; Baek, J.; Rosen, B.D.; Tar’an, B.; et al. Draft genome sequence of chickpea (Cicer arietinum) provides a resource for trait improvement. Nat. Biotechnol. 2013, 31, 240–246. [Google Scholar] [CrossRef] [Green Version]
  219. Jain, D.; Chattopadhyay, D. Analysis of gene expression in response to water deficit of chickpea (Cicer arietinum L.) varieties differing in drought tolerance. BMC Plant Biol. 2010, 10, 24. [Google Scholar] [CrossRef]
  220. Roorkiwal, M.; Sharma, P.C. Sequence similarity-based identification of abiotic stress responsive genes in chickpea. Bioinformation 2012, 8, 92–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Varshney, R.K.; Mir, R.R.; Bhatia, S.; Thudi, M.; Hu, Y.; Azam, S.; Zhang, Y.; Jagannathan, D. Integrated physical map with the genetic maps and reference genome sequence for chickpea (Cicer arietinum L.) improvement. Funct. Integr. Genom. 2014, 14, 59–73. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Muriuki, R.; Kimurto, P.K.; Towett, B.K.; Vadez, V.; Gangarao, R. Study of root traits of chickpea (Cicer arietinum L.) under drought stress. Afr. J. Plant Sci. 2020, 14, 420–435. [Google Scholar]
  223. Upadhyaya, H.D.; Dwivedi, S.L.; Baum, M.; Varshney, R.K.; Udupa, S.M.; Gowda, C.L.L.; Hoisington, D.; Singh, S. Genetic structure, diversity, and allelic richness in composite collection and reference set in chickpea (Cicer arietinum L.). BMC Plant Biol. 2008, 8, 106. [Google Scholar] [CrossRef] [Green Version]
  224. Roorkiwal, M.; Nayak, S.N.; Thudi, M.; Upadhyaya, H.D.; Brunel, D.; Mournet, P.; This, D.; Sharma, P.C.; Varshney, R.K. Allele diversity for abiotic stress responsive candidate genes in chickpea reference set using gene-based SNP markers. Front. Plant Sci. 2014, 5, 248. [Google Scholar] [CrossRef]
  225. Roorkiwal, M.; Bharadwaj, C.; Barmukh, R.; Dixit, G.P.; Thudi, M.; Gaur, P.M.; Chaturvedi, S.K.; Fikre, A.; Hamwieh, A.; Kumar, S.; et al. Integrating genomics for chickpea improvement: Achievements and opportunities. Theor. Appl. Genet. 2020, 133, 1703–1720. [Google Scholar] [CrossRef] [Green Version]
  226. Upadhyaya, H.D.; Bajaj, D.; Das, S.; Kumar, V.; Gowda, C.L.L.; Sharma, S.; Tyagi, A.K.; Parida, S.K. Genetic dissection of seed-iron and zinc concentrations in chickpea. Sci. Rep. 2016, 6, 24050. [Google Scholar] [CrossRef] [Green Version]
  227. Zhao, K.; Tung, C.W.; Eizenga, G.C.; Wright, M.H.; Ali, M.L.; Price, A.H.; Norton, G.J.; Islam, M.R.; Reynolds, A.; Mezey, J.; et al. Genome wide association mapping reveals a rich genetic architecture of complex traits in Oryza sativa. Nat. Commun. 2011, 2, 467. [Google Scholar] [CrossRef] [Green Version]
  228. Pasam, R.K.; Sharma, R.; Malosetti, M.; van Eeuwijk, F.A.; Haseneyer, G.; Kilian, B.; Graner, A. Genome-wide association studies for agronomical traits in a worldwide spring barley collection. BMC Plant Biol. 2012, 12, 16. [Google Scholar] [CrossRef] [Green Version]
  229. Morris, G.P.; Ramu, P.; Deshpande, S.P.; Hash, C.T.; Shah, T.; Upadhyaya, H.D.; Riera-Lizarazu, O.; Brown, P.J.; Acharya, C.B.; Mitchell, S.E.; et al. Population genomic and genome-wide association studies of agroclimatic traits in sorghum. Proc. Natl. Acad. Sci. USA 2013, 110, 453–458. [Google Scholar] [CrossRef] [Green Version]
  230. Karaca, N.; Ates, D.; Nemli, S.; Ozkuru, E.; Yilmaz, H.; Yagmur, B.; Kartal, C.; Tosun, M.; Ozdestan, O.; Otles, S.; et al. Identification of SNP Markers Associated with Iron and Zinc Concentrations in Cicer Seeds. Curr. Genom. 2020, 21, 212–223. [Google Scholar] [CrossRef] [PubMed]
  231. Jha, U.; Jha, R.; Bohra, A.; Manjunatha, L.; Saabale, P.; Parida, S.; Singh, N. Association mapping of genomic loci linked with Fusarium wilt resistance (Foc2) in chickpea. Plant Genet. Resour. Charact. Util. 2021, 19, 195–202. [Google Scholar] [CrossRef]
  232. Varshney, R.K.; Pandey, M.K.; Bohra, A.; Singh, V.K.; Thudi, M.; Saxena, R.K. Toward the sequence-based breeding in legumes in the post-genome sequencing era. Theor. Appl. Genet. 2019, 132, 797–816. [Google Scholar] [CrossRef] [Green Version]
  233. Samineni, S.; Sajja, S.B.; Mondal, B.; Chand, U.; Thudi, M.; Varshney, R.K.; Gaur, P.M. MAGIC lines in chickpea: Development and exploitation of genetic diversity. Euphytica (TSI) 2021, 217, 1–12. [Google Scholar] [CrossRef]
  234. Gangurde, S.S.; Wang, H.; Yaduru, S.; Pandey, M.K.; Fountain, J.C.; Chu, Y.; Isleib, T.; Holbrook, C.C.; Xavier, A.; Culbreath, A.K.; et al. Nested-association mapping (NAM)-based genetic dissection uncovers candidate genes for seed and pod weights in peanut (Arachis hypogaea). Plant Biotechnol. J. 2020, 18, 1457–1471. [Google Scholar] [CrossRef] [Green Version]
  235. McMullen, M.D.; Kresovich, S.; Villeda, H.S.; Bradbury, P.; Li, H.; Sun, Q.; Flint-Garcia, S.; Thornsberry, J.; Acharya, C.; Bottoms, C.; et al. Genetic properties of the maize nested association mapping population. Science 2009, 325, 737–740. [Google Scholar] [CrossRef] [Green Version]
  236. Bajaj, D.; Srivastava, R.; Nath, M.; Tripathi, S.; Bharadwaj, C.; Upadhyaya, H.D.; Tyagi, A.K.; Parida, S.K. EcoTILLING-Based Association Mapping Efficiently Delineates Functionally Relevant Natural Allelic Variants of Candidate Genes Governing Agronomic Traits in Chickpea. Front. Plant Sci. 2016, 7, 450. [Google Scholar] [CrossRef] [Green Version]
  237. Palakurthi, R.; Jayalakshmi, V.; Kumar, Y.; Kulwal, P.; Yasin, M.; Kute, N.S.; Laxuman, C.; Yeri, S.; Vemula, A.; Rathore, A.; et al. Translational Chickpea Genomics Consortium to Accelerate Genetic Gains in Chickpea (Cicer arietinum L.). Plants 2021, 10, 2583. [Google Scholar] [CrossRef]
  238. Varshney, R.K.; Saxena, R.K.; Upadhyaya, H.D.; Khan, A.W.; Yu, Y.; Kim, C.; Rathore, A.; Kim, D.; Kim, J.; An, S.; et al. Whole-genome resequencing of 292 pigeonpea accessions identifes genomic regions associated with domestication and agronomic traits. Nat. Genet. 2017, 49, 1082–1088. [Google Scholar] [CrossRef]
  239. Arriagada, O.; Cacciuttolo, F.; Cabeza, R.A.; Carrasco, B.; Schwember, A.R. A Comprehensive Review on Chickpea (Cicer arietinum L.) Breeding for Abiotic Stress Tolerance and Climate Change Resilience. Int. J. Mol. Sci. 2022, 23, 6794. [Google Scholar] [CrossRef]
  240. Kanzi, A.M.; San, J.E.; Chimukangara, B.; Wilkinson, E.; Fish, M.; Ramsuran, V.; de Oliveira, T. Next Generation Sequencing and Bioinformatics Analysis of Family Genetic Inheritance. Front. Genet. 2020, 11, 544162. [Google Scholar] [CrossRef] [PubMed]
  241. Kosgei, A.J.; PKKimurtoPMGaurMAYeboahSKOffei, E.Y. Danquah, Introgression of drought tolerance root traits into Kenyan commercial chickpea varieties using marker assisted backcrossing. Afr. Crop. Sci. J. 2022, 30, 31–50. [Google Scholar] [CrossRef]
  242. Jain, A.; Govindaraj, G.M.; Edavazhippurath, A.; Faisal, N.; Bhoyar, R.C.; Gupta, V.; Uppuluri, R.; Manakkad, S.P.; Kashyap, A.; Kumar, A.; et al. Whole genome sequencing identifies novel structural variant in a large Indian family affected with X-linked agammaglobulinemia. PLoS ONE 2021, 16, e0254407. [Google Scholar] [CrossRef] [PubMed]
  243. Jaganathan, D.; Thudi, M.; Kale, S.; Azam, S.; Roorkiwal, M.; Gaur, P.M.; Kishor, P.B.; Nguyen, H.; Sutton, T.; Varshney, R.K. Genotyping-by-sequencing based intra-specific genetic map refines a ‘‘QTL-hotspot” region for drought tolerance in chickpea. Mol. Genet. Genom. 2015, 290, 559–571. [Google Scholar] [CrossRef]
  244. Thudi, M.; Chitikineni, A.; Liu, X.; He, W.; Roorkiwal, M.; Yang, W.; Jian, J.; Doddamani, D.; Gaur, P.M.; Rathore, A.; et al. Recent breeding programs enhanced genetic diversity in both desi and kabuli varieties of chickpea (Cicer arietinum L.). Sci. Rep. 2016, 6, 38636. [Google Scholar] [CrossRef] [Green Version]
  245. Varshney, R.; Thudi, M.; Upadhyaya, H.; Dwivedi, S.; Udupa, S.; Furman, B.; Baum, M.; Hoisington, D. A SSR kit to study genetic diversity in chickpea (Cicer arietinum L.). Plant Genet. Resour. 2014, 12, S118–S120. [Google Scholar] [CrossRef]
  246. Roorkiwal, M.; Jain, A.; Kale, S.M.; Doddamani, D.; Chitikineni, A.; Thudi, M.; Varshney, R.K. Development and evaluation of highdensityAxiom® CicerSNP Array for high-resolution genetic mapping and breeding applications in chickpea. Plant Biotechnol. J. 2018, 16, 890–901. [Google Scholar] [CrossRef] [Green Version]
  247. Rajkumar, M.S.; Garg, R.; Jain, M. Genome resequencing reveals DNA polymorphisms associated with seed size/weight determination in chickpea. Genomics 2021, 113, 1458–1468. [Google Scholar] [CrossRef]
  248. Varshney, R.K.; Bohra, A.; Yu, J.; Graner, A.; Zhang, Q.; Sorrells, M.E. Designing Future Crops: Genomics-Assisted Breeding Comes of Age. Trends Plant Sci. 2021, 26, 631–649. [Google Scholar] [CrossRef]
  249. Varshney, R.K.; Graner, A.; Sorrells, M.E. Genic microsatellite markers in plants: Features and applications. Trends Biotechnol. 2005, 23, 48–55. [Google Scholar] [CrossRef]
  250. Varshney, R.K.; Roorkiwal, M.; Sun, S.; Bajaj, P.; Chitikineni, A.; Thudi, M.; Singh, N.P.; Du, X.; Upadhyaya, H.D.; Khan, A.W.; et al. A chickpea genetic variation map based on the sequencing of 3366 genomes. Nature 2021, 599, 622–627. [Google Scholar] [CrossRef] [PubMed]
  251. Coletta, D.R.; Qiu, Y.; Ou, S.; Hufford, M.B.; Hirsch, C.N. How the pan-genome is changing crop genomics and improvement. Genome Biol. 2021, 22, 3. [Google Scholar] [CrossRef]
  252. Hardison, R.C. Comparative genomics. PLoS Biol. 2003, 1, E58. [Google Scholar] [CrossRef] [PubMed]
  253. Boutte, J.; Maillet, L.; Chaussepied, T.; Letort, S.; Aury, J.M.; Belser, C.; Boideau, F.; Brunet, A.; Coriton, O.; Deniot, G.; et al. Genome Size Variation and Comparative Genomics Reveal Intraspecific Diversity in Brassica rapa. Front. Plant Sci. 2020, 11, 577536. [Google Scholar] [CrossRef] [PubMed]
  254. Ebler, J.; Ebert, P.; Clarke, W.E.; Rausch, T.; Audano, P.A.; Houwaart, T.; Mao, Y.; Korbel, J.O.; Eichler, E.E.; Zody, M.C.; et al. Pangenome-based genome inference allows efficient and accurate genotyping across a wide spectrum of variant classes. Nat. Genet. 2022, 54, 518–525. [Google Scholar] [CrossRef] [PubMed]
  255. Khan, A.W.; Garg, V.; Roorkiwal, M.; Golicz, A.A.; Edwards, D.; Varshney, R.K. Super-Pangenome by Integrating the Wild Side of a Species for Accelerated Crop Improvement. Trends Plant Sci. 2020, 25, 148–158. [Google Scholar] [CrossRef] [Green Version]
  256. Lei, L.; Goltsman, E.; Goodstein, D.; Wu, G.A.; Rokhsar, D.S.; Vogel, P.J. Plant Pan-Genomics Comes of Age. Annu. Rev. Plant Biol. 2021, 72, 411–435. [Google Scholar] [CrossRef]
  257. Tay Fernandez, C.G.; Nestor, B.J.; Danilevicz, M.F.; Gill, M.; Petereit, J.; Bayer, P.E.; Finnegan, P.M.; Batley, J.; Edwards, D. Pangenomes as a Resource to Accelerate Breeding of Under-Utilised Crop Species. Int. J. Mol. Sci. 2022, 23, 2671. [Google Scholar] [CrossRef]
  258. Jha, U.C.; Bohra, A.; Nayyar, H. Advances in “omics” approaches to tackle drought stress in grain legumes. Plant Breed. 2019, 139, 1–27. [Google Scholar] [CrossRef] [Green Version]
  259. Yang, Y.; Saand, M.A.; Huang, L.; Abdelaal, W.B.; Zhang, J.; Wu, Y.; Li, J.; Sirohi, M.H.; Wang, F. Applications of Multi-Omics Technologies for Crop Improvement. Front. Plant Sci. 2021, 12, 563953. [Google Scholar] [CrossRef]
  260. Kumar, R.; Sharma, V.; Suresh, S.; Ramrao, D.P.; Veershetty, A.; Kumar, S.; Priscilla, K.; Hangargi, B.; Narasanna, R.; Pandey, M.K.; et al. Understanding Omics Driven Plant Improvement and de novo Crop Domestication: Some Examples. Front. Genet. 2021, 12, 637141. [Google Scholar] [CrossRef] [PubMed]
  261. Sreenivasulu, N.; Kishor, P.B.K.; Varshney, R.K.; Altschmied, L. Mining functional information from cereal genomics–the utility of expressed sequence tags. Curr. Sci. 2002, 83, 965–973. [Google Scholar]
  262. Pradhan, S.; Bandhiwal, N.; Shah, N.; Kant, C.; Gaur, R.; Bhatia, S. Global transcriptome analysis of developing chickpea (Cicer arietinum L.) seeds. Front. Plant Sci. 2014, 5, 698. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Agarwal, G.; Jhanwar, S.; Priya, P.; Singh, V.K.; Saxena, M.S.; Parida, S.K.; Garg, R.; Tyagi, A.K.; Jain, M. Comparative Analysis of Kabuli Chickpea Transcriptome with Desi and Wild Chickpea Provides a Rich Resource for Development of Functional Markers. PLoS ONE 2012, 7, e52443. [Google Scholar] [CrossRef] [Green Version]
  264. Garg, R.; Patel, R.K.; Jhanwar, S.; Priya, P.; Bhattacharjee, A.; Yadav, G.; Bhatia, S.; Chattopadhyay, D.; Tyagi, A.K.; Jain , M. Gene discovery and tissue-specific transcriptome analysis in chickpea with massively parallel pyrosequencing and web resource development. Plant Physiol. 2011, 1564, 1661–1678. [Google Scholar] [CrossRef] [Green Version]
  265. Kumar, N.; Singh, S.S.; Ghosh, P.K.; Hazra, K.K.; Venkatesh, M.S.; Praharaj, C.S.; Singh, M.K.; Kumar MSenthilBasu, P.S.; Yadav, A.; Yadav, S.L.; et al. Improving chickpea productivity in rice-fallow of Indo-Gangetic Plain with soil moisture conservation and cultivar selection. J. Food Legumes 2020, 33, 28–35. [Google Scholar]
  266. Kaashyap, M.; Ford, R.; Mann, A.; Varshney, R.K.; Siddique, K.H.M.; Mantri, N. Comparative Flower Transcriptome Network Analysis Reveals DEGs Involved in Chickpea Reproductive Success during Salinity. Plants 2022, 11, 434. [Google Scholar] [CrossRef]
  267. Molina, C.; Rotter, B.; Horres, R.; Udupa, S.M.; Besser, B.; Bellarmino, L.; Baum, M.; Matsumura, H.; Terauchi, R.; Kahl, G.; et al. SuperSAGE: The drought stress-responsive transcriptome of chickpea roots. BMC Genom. 2008, 9, 553. [Google Scholar] [CrossRef] [Green Version]
  268. Carrari, F.; Fernie, A.R.; Iusem, N.D. Heard it through the grapevine? ABA and sugar cross-talk: The ASR story. Trends Plant Sci. 2004, 9, 57–59. [Google Scholar] [CrossRef]
  269. Kumar, M.; Kumar Patel, M.; Kumar, N.; Bajpai, A.B.; Siddique, K.H.M. Metabolomics and Molecular Approaches Reveal Drought Stress Tolerance in Plants. Int. J. Mol. Sci. 2021, 22, 9108. [Google Scholar] [CrossRef]
  270. Deokar, A.A.; Kondawar, V.; Jain, P.K.; Karuppayil, S.M.; Raju, N.L.; Vadez, V.; Varshney, R.K.; Srinivasan, R. Comparative analysis of expressed sequence tags (ESTs) between drought-tolerant and -susceptible genotypes of chickpea under terminal drought stress. BMC Plant Biol. 2011, 11, 70. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  271. Garg, R.; Shankar, R.; Thakkar, B.; Kudapa, H.; Krishnamurthy, L.; Mantri, N.; Varshney, R.K.; Bhatia, S.; Jain, M. Transcriptome analyses reveal genotype- and developmental stage-specific molecular responses to drought and salinity stresses in chickpea. Sci. Rep. 2016, 6, 19228. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  272. Raes, J.; Rohde, A.; Christensen, J.H.; Van de Peer, Y.; Boerjan, W. Genome-wide characterization of the lignification toolbox in Arabidopsis. Plant Physiol. 2003, 133, 1051–1071. [Google Scholar] [CrossRef] [PubMed]
  273. Kumar, N.; Bharadwaj, C.; Soni, A.; Sachdeva, S.; Yadav, M.C.; Pal, M.; Soren, K.R.; Meena, M.C.; Roorkiwal, M.; Varshney, R.K.; et al. Physio-morphological and molecular analysis for salt tolerance in chickpea (Cicer arietinum). Indian J. Agric. Sci. 2020, 90, 804–808. [Google Scholar]
  274. Ahmad, B.; Azeem, F.; Ali, M.A.; Nawaz, M.A.; Nadeem, H.; Abbas, A.; Batool, R.; Atif, R.M.; Ijaz, U.; Nieves-Cordones, M.; et al. Genome-wide identification and expression analysis of two component system genes in Cicer arietinum. Genomics 2020, 112, 1371–1383. [Google Scholar] [CrossRef] [PubMed]
  275. Mashaki, K.M.; Garg, V.; Ghomi, A.A.N.; Kudapa, H.; Chitikineni, A.; Nezhad, K.Z.; Yamchi, A.; Soltanloo, H.; Varshney, R.K.; Thudi, M. RNA-Seq analysis revealed genes associated with drought stress response in kabuli chickpea (Cicer arietinum L.). PLoS ONE 2018, 13, e0199774. [Google Scholar] [CrossRef] [Green Version]
  276. Khan, N.; Bano, A.; Rahman, M.A.; Rathinasabapathi, B.; Babar, M.A. UPLC-HRMS-based untargeted metabolic profiling reveals changes in chickpea (Cicer arietinum) metabolome following long-term drought stress. Plant Cell Environ. 2019, 42, 115–132. [Google Scholar] [CrossRef] [Green Version]
  277. Dias, D.A.; Hill, C.B.; Jayasinghe, N.S.; Atieno, J.; Sutton, T.; Roessner, U. Quantitative profling of polar primary metabolites of two chickpea cultivars with contrasting responses to salinity. J. Chromatogr. B Anal. Technol. Biomed. Life Sci 2015, 1000, 1. [Google Scholar] [CrossRef] [Green Version]
  278. Shariatipour, N.; Heidari, B. Investigation of Drought and Salinity Tolerance Related Genes and their Regulatory Mechanisms in Arabidopsis (Arabidopsis thaliana). Open Bioinform. J. 2018, 11, 12–28. [Google Scholar] [CrossRef]
  279. Bhushan, D.; Pandey, A.; Choudhary, M.K.; Datta, A.; Chakraborty, S.; Chakraborty, N. Comparative proteomics analysis of differentially expressed proteins in chickpea extracellular matrix during dehydration stress. Mol. Cell Prot. 2007, 6, 1868–1884. [Google Scholar] [CrossRef] [Green Version]
  280. Jaiswal, D.; Mishra, P.; Subba, P.; Rathi, D.; Chakraborty, S.; Chakraborty, N. Membrane-associated proteomics of chickpea identifies Sad1/UNC-84 protein (CaSUN1), a novel component of dehydration signaling. Sci. Rep. 2014, 4, 4177. [Google Scholar] [CrossRef] [Green Version]
  281. Vessal, S.; Arefian, M.; Siddique, K.H.M. Proteomic responses to progressive dehydration stress in leaves of chickpea seedlings. BMC Genom. 2020, 21, 523. [Google Scholar] [CrossRef] [PubMed]
  282. Gupta, S.; Mishra, S.K.; Misra, S.; Pandey, V.; Agrawal, L.; Nautiyal, C.S.; Chauhan, P.S. Revealing the complexity of protein abundance in chickpea root under drought-stress using a comparative proteomics approach. Plant Physiol. Biochem. 2020, 151, 88–102. [Google Scholar] [CrossRef] [PubMed]
  283. Azimi, S.; Kaur, T.; Gandhi, T.K. Water Stress Identification in Chickpea Plant Shoot Images using Deep Learning. In Proceedings of the 2020 IEEE 17th India Council International Conference (INDICON), New Delhi, India, 10–13 December 2020; pp. 1–7. [Google Scholar] [CrossRef]
  284. Clark, R.T.; MacCurdy, R.B.; Jung, J.K.; Shaff, J.E.; McCouch, S.R.; Aneshansley, D.J.; Kochian, L.V. Three-dimensional root phenotyping with a novel imaging and software platform. Plant Physiol. 2011, 156, 455–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Zhu, M.; Zhao, S. Candidate gene identification approach: Progress and challenges. Int. J. Biol. Sci. 2007, 3, 420–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Cano-Gamez, E.; Trynka, G. From GWAS to Function: Using Functional Genomics to Identify the Mechanisms Underlying Complex Diseases. Front. Genet. 2020, 11, 424. [Google Scholar] [CrossRef]
  287. Zhong, C.; Sun, S.; Li, Y.; Duan, C.; Zhu, Z. Next-generation sequencing to identify candidate genes and develop diagnostic markers for a novel Phytophthora resistance gene, RpsHC18, in soybean. Theor. Appl. Genet. 2018, 131, 525–538. [Google Scholar] [CrossRef]
  288. Dossa, K.; Li, D.; Zhou, R.; Yu, J.; Wang, L.; Zhang, Y.; You, J.; Liu, A.; Mmadi, M.A.; Fonceka, D.; et al. The genetic basis of drought tolerance in the high oil crop Sesamum indicum. Plant Biotechnol. J. 2019, 17, 1788–1803. [Google Scholar] [CrossRef] [Green Version]
  289. Bhattarai, G.; Shi, A.; Feng, C.; Dhillon, B.; Mou, B.; Correll, J.C. Genome Wide Association Studies in Multiple Spinach Breeding Populations Refine Downy Mildew Race 13 Resistance Genes. Front. Plant Sci. 2020, 11, 563187. [Google Scholar] [CrossRef]
  290. Das, A.; Basu, P.S.; Kumar, M.; Ansari, J.; Shukla, A.; Thakur, S.; Singh, P.; Datta, S.; Chaturvedi, S.K.; Sheshshayee, M.S.; et al. Transgenic chickpea (Cicer arietinum L.) harbouring AtDREB1a are physiologically better adapted to water deficit. BMC Plant Biol. 2021, 21, 39. [Google Scholar] [CrossRef]
  291. Puhakainen, T.; Hess, M.W.; Makela, P.; Svensson, J.; Heino, P.; Palva, E.T. Over expression of multiple dehydrin genes enhances tolerance to freezing stress in Arabidopsis. Plant Mol. Biol. 2004, 54, 743–753. [Google Scholar] [CrossRef] [PubMed]
  292. Allagulova, C.R.; Gimalov, F.R.; Shakirova, F.M.; Vakhitov, V.A. The plant dehydrins: Structure and putative functions. Biochemistry 2003, 68, 1157–1165. [Google Scholar] [CrossRef]
  293. Kumar, T.; Tiwari, N.; Bharadwaj, C.; Sarker, A.; Pappula, S.P.R.; Singh, S.; Singh, M. Identification of Allelic Variation in Drought Responsive Dehydrin Gene Based on Sequence Similarity in Chickpea (Cicer arietinum L.). Front. Genet. 2020, 11, 584527. [Google Scholar] [CrossRef] [PubMed]
  294. Ichimura, K.; Mizoguchi, T.; Yoshida, R.; Yuasa, T.; Shinozaki, K. Various abiotic stresses rapidly activate Arabidopsis MAP kinases AtMPK4 and AtMPK6. Plant J. 2000, 24, 655–665. [Google Scholar] [CrossRef]
  295. Stiti, N.; Missihoun, T.D.; Kotchoni, S.O.; Kirch, H.H.; Bartels, D. Aldehyde dehydrogenases in Arabidopsis thaliana: Biochemical requirements, metabolic pathways, and functional analysis. Front. Plant Sci. 2011, 2, 65. [Google Scholar] [CrossRef] [Green Version]
  296. Park, J.M.; Park, C.J.; Lee, S.B.; Ham, B.K.; Shin, R.; Paek, K.H. Overexpression of the tobacco Tsi1 gene encoding an EREBP/AP2-type transcription factor enhances resistance against pathogen attack and osmotic stress in tobacco. Plant Cell 2001, 13, 1035–1046. [Google Scholar] [CrossRef] [Green Version]
  297. Shukla, R.K.; Tripathi, V.; Jain, D.; Yadav, R.K.; Chattopadhyay, D. CAP2 enhances germination of transgenic tobacco seeds at high temperature and promotes heat stress tolerance in yeast. FEBS J. 2009, 276, 5252–5262. [Google Scholar] [CrossRef]
  298. Shukla, U.C.; Yadav, O.P. Effect of phosphorus and zinc on nodulation and nitrogen fixation in chickpea (Cicer arietinum L.). J. Plant Soil 2004, 65, 239–248. [Google Scholar] [CrossRef]
  299. Gu, H.; Jia, Y.; Wang, X.; Chen, Q.; Shi, S.; Ma, L.; Zhang, J.; Zhang, H.; Ma, H. Identification and characterization of a LEA family gene CarLEA4 from chickpea (Cicer arietinum L.). Mol. Biol. Rep. 2012, 39, 3565–3572. [Google Scholar] [CrossRef]
  300. Sharma, K.D.; Patil, D.; Kiran, A. Characterization and differential expression of sucrose and starch metabolism genes in contrasting chickpea (Cicer arietinum L.) genotypes under low temperature. J. Genet. 2021, 100, 71. [Google Scholar] [CrossRef]
  301. Caballo, C.; Berbel, A.; Ortega, R.; Gil, J.; Millán, T.; Rubio, J.; Madueño, F. The SINGLE FLOWER (SFL) gene encodes a MYB transcription factor that regulates the number of flowers produced by the inflorescence of chickpea. New Phytol. 2022, 234, 827–836. [Google Scholar] [CrossRef] [PubMed]
  302. Yang, C.Y.; Chen, Y.C.; Jauh, G.Y.; Wang, C.S. A Lily ASR protein involves abscisic acid signaling and confers drought and salt resistance in Arabidopsis. Plant Physiol. 2005, 139, 836–846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Li, W.; Huang, D.; Wang, B.; Hou, X.; Zhang, R.; Yan, M.; Liao, W. Changes of starch and sucrose content and related gene expression during the growth and development of Lanzhou lily bulb. PLoS ONE 2022, 17, e0262506. [Google Scholar] [CrossRef]
  304. Meenakshi Kumar, A.; Kumar, V.; Dubey, A.K.; Narayan, S.; Sawant, M.V.; Pande, V.; Sirke, P.; Sanyal, I. CAMTA transcription factor enhances salinity and drought tolerance in chickpea (Cicer arietinum L.). Plant Cell Tissue Organ Cult. 2022, 148, 319–330. [Google Scholar] [CrossRef]
  305. Yu, X.; Liu, Y.; Wang, S.; Tao, Y.; Wang, Z.; Shu, Y.; Peng, H.; Mijiti, A.; Wang, Z.; Zhang, H.; et al. CarNAC4, a NAC-type chickpea transcription factor conferring enhanced drought and salt stress tolerances in Arabidopsis. Plant Cell Rep. 2016, 35, 613–627. [Google Scholar] [CrossRef] [PubMed]
  306. Nguyen, K.H.; Ha, C.V.; Watanabe, Y.; Tran, U.T.; Nasr Esfahani, M.; Nguyen, D.V.; Tran, L.-S.P. Correlation between differential drought tolerability of two contrasting drought-responsive chickpea cultivars and differential expression of a subset of CaNAC genes under normal and dehydration conditions. Front. Plant Sci. 2015, 6, 449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. La, H.V.; Chu, H.D.; Tran, C.D.; Nguyen, K.H.; Le, Q.T.N.; Hoang, C.M.; Cao, B.P.; Pham, A.T.C.; Nguyen, B.D.; Nguyen, T.Q.; et al. Insights into the gene and protein structures of the CaSWEET family members in chickpea (Cicer arietinum), and their gene expression patterns in different organs under various stress and abscisic acid treatments. Gene 2022, 819, 146210. [Google Scholar] [CrossRef]
  308. Agarwal, P.K.; Jha, B. Transcription factors in plants and ABA dependent and independent abiotic stress signalling. Biol. Plant 2010, 54, 201–212. [Google Scholar] [CrossRef]
  309. Riechmann, J.L.; Heard, J.; Martin, G.; Reuber, L.; Jiang, C.; Keddie, J.; Adam, L.; Pineda, O.; Ratcliffe, O.J.; Samaha, R.R.; et al. Arabidopsis transcription factor: Genome wide comparative analysis among eukaryotes. Science 2000, 290, 2105–2110. [Google Scholar] [CrossRef]
  310. Lata, C.; Prasad, M. Role of DREBs in regulation of abiotic stress responses in plants. J. Exp. Bot. 2011, 62, 4731–4748. [Google Scholar] [CrossRef] [Green Version]
  311. Anbazhagan, K.; Bhatnagar-Mathur, P.; Vadez, V.; Dumbala, S.R.; Kishor, P.B.; Sharma, K.K. DREB1A overexpression in transgenic chickpea alters key traits influencing plant water budget across water regimes. Plant Cell Rep. 2015, 34, 199–210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  312. Hanin, M.; Brini, F.; Ebel, C.; Toda, Y.; Takeda, S.; Masmoudi, K. Plant dehydrins and stress tolerance: Versatile proteins for complex mechanisms. Plant Signal. Behav. 2011, 6, 1503–1509. [Google Scholar] [CrossRef] [PubMed]
  313. Wise, M.J.; Tunnacliffe, A. POPP the question: What do LEA protein do? Trends Plant Sci. 2004, 9, 13–17. [Google Scholar] [CrossRef] [PubMed]
  314. Puhakainen, T.; Li, C.; Boije-Malm, M.; Kangasjarvi, J.; Heino, P.; Palva, E.T. Short-day potentiation of low temperature-induced gene expression of a C-repeat-binding factor-controlled gene during cold acclimation in silver birch. Plant Physiol. 2004, 136, 4299–4307. [Google Scholar] [CrossRef] [Green Version]
  315. Rudrabhatla, P.; Reddy, M.M.; Rajasekharan, R. Genome-wide analysis and experimentation of plant serine/threonine/tyrosine-specific protein kinases. Plant Mol. Biol. 2006, 60, 293–319. [Google Scholar] [CrossRef]
  316. Petrivalský, M.; Brauner, F.; Luhová, L.; Gagneul, D.; Sebela, M. Aminoaldehyde dehydrogenase activity during wound healing of mechanically injured pea seedlings. J. Plant Physiol. 2007, 164, 1410–1418. [Google Scholar] [CrossRef]
  317. Skibbe, D.S.; Liu, F.; Wen, T.J.; Yandeau, M.D.; Cui, X.; Cao, J.; Simmons, C.R.; Schnable, P.S. Characterization of the aldehyde dehydrogenase gene families of Zea mays and Arabidopsis. Plant Mol. Biol. 2002, 48, 751–764. [Google Scholar] [CrossRef]
  318. Shpak, E.D.; Berthiaume, C.T.; Hill, E.J.; Torii, K.U. Synergistic interaction of three ERECTA-family receptor-like kinases controls Arabidopsis organ growth and flower development by promoting cell proliferation. Development 2004, 131, 1491–1501. [Google Scholar] [CrossRef] [Green Version]
  319. Mandel, T.; Moreau, F.; Kutsher, Y.; Fletcher, J.C.; Carles, C.C.; Eshed Williams, L. The ERECTA receptor kinase regulates Arabidopsis shoot apical meristem size, phyllotaxy and floral meristem identity. Development 2014, 141, 830–841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Zhu, S.; Shi, W.; Jie, Y.; Zhou, Q.; Song, C. A MYB transcription factor, BnMYB2, cloned from ramie (Boehmeria nivea) is involved in cadmium tolerance and accumulation. PLoS ONE 2020, 15, e0233375. [Google Scholar] [CrossRef]
  321. Wei, Q.; Chen, R.; Wei, X.; Liu, Y.; Zhao, S.; Yin, X.; Xie, T. Genome-wide identification of R2R3-MYB family in wheat and functional characteristics of the abiotic stress responsive gene TaMYB344. BMC Genom. 2020, 21, 792. [Google Scholar] [CrossRef] [PubMed]
  322. Islam, K.; Rawoof, A.; Ahmad, I.; Dubey, M.; Momo, J.; Ramchiary, N. Capsicum chinense MYB Transcription Factor Genes: Identification, Expression Analysis, and Their Conservation and Diversification with Other Solanaceae Genomes. Front. Plant Sci. 2021, 12, 721265. [Google Scholar] [CrossRef] [PubMed]
  323. Heidari, P.; Mazloomi, F.; Nussbaumer, T.; Barcaccia, G. Insights into the SAM Synthetase Gene Family and Its Roles in Tomato Seedlings under Abiotic Stresses and Hormone Treatments. Plants 2020, 9, 586. [Google Scholar] [CrossRef]
  324. Sun, Y.; Locasale, J.W. Rethinking the bioavailability and cellular transport properties of S-adenosylmethionine. Cell Stress 2021, 6, 1–5. [Google Scholar] [CrossRef] [PubMed]
  325. Radadiya, N.; Parekh, V.P.; Dobariya, B.; Mahatma, L.; Mahatma, M.K. Abiotic stresses alter expression of S-Adenosylmethionine synthetase gene, polyamines and antioxidant activity in pigeon pea (Cajanus cajan L.). Legume Res. 2016, 2016, 905–913. [Google Scholar]
  326. Joo, J.; Lee, Y.H.; Kim, Y.K.; Nahm, B.H.; Song, S.I. Abiotic stress responsive rice ASR1 and ASR3 exhibit different tissue-dependent sugar and hormone-sensitivities. Mol. Cells 2013, 35, 421–435. [Google Scholar] [CrossRef] [Green Version]
  327. Philippe, R.; Courtois, B.; McNally, K.L.; Mournet, P.; El-Malki, R.; Paslier, M.C.; Fabre, D.; Billot, C.; Brunel, D.; Glaszmann, J.C.; et al. Structure, allelic diversity and selection of Asr genes, candidate for drought tolerance, in Oryza sativa L. and wild relatives. Theor. Appl. Genet. 2010, 121, 769–787. [Google Scholar] [CrossRef]
  328. Cortés, A.J.; Monserrate, F.A.; Ramírez-Villegas, J.; Madriñán, S.; Blair, M.W. Drought tolerance in wild plant populations: The case of common beans (Phaseolus vulgaris L.). PLoS ONE 2013, 8, e62898. [Google Scholar] [CrossRef] [Green Version]
  329. Yoshida, T.; Fujita, Y.; Sayama, H.; Kidokoro, S.; Maruyama, K.; Mizoi, J.; Shinozaki, K.; Yamaguchi-Shinozaki, K. AREB1, AREB2 and AFB2 are master transcription factors that cooperatively regulate ABRE dependent ABA signalling involved in drought stress tolerance and require ABA for full activation. Plant J. 2010, 61, 672–685. [Google Scholar] [CrossRef]
  330. Stein, O.; Granot, D. An Overview of Sucrose Synthases in Plants. Front. Plant Sci. 2019, 10, 95. [Google Scholar] [CrossRef] [Green Version]
  331. Yao, D.; Gonzales-Vigil, E.; Mansfield, S.D. Arabidopsis sucrose synthase localization indicates a primary role in sucrose translocation in phloem. J. Exp. Bot. 2020, 71, 1858–1869. [Google Scholar] [CrossRef] [PubMed]
  332. Kumar, A.; Turner, N. Growth and sucrose synthase activity of developing chickpea (Cicer arietinum L.) seeds under field conditions. Aust. J. Crop. Sci. 2009, 3, 20–27. [Google Scholar]
  333. Okamuro, J.K.; Caster, B.; Villarroel, R.; Montagu, M.V.; Jofuku, K.D. The AP2 domain of APETALA2 defines a large new family of DNA binding proteins in Arabidopsis. Proc. Natl. Acad. Sci. USA 1997, 94, 7076–7081. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  334. Shukla, R.K.; Raha, S.; Tripathi, V.; Chattopadhyay, D. Expression of CAP2, an APETALA2-family transcription factor from chickpea, enhances growth and tolerance to dehydration and salt stress in transgenic tobacco. Plant Physiol. 2006, 142, 113–123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  335. Jain, D.; Chattopadhyay, D. Promoter of CaZF, a Chickpea Gene That Positively Regulates Growth and Stress Tolerance, Is Activated by an AP2-Family Transcription Factor CAP2. PLoS ONE 2013, 8, e56737. [Google Scholar] [CrossRef] [PubMed]
  336. Nemati, F.; Ghanati, F.; Ahmadi Gavlighi, H.; Sharifi, M. Comparison of sucrose metabolism in wheat seedlings during drought stress and subsequent recovery. Biol. Plant 2018, 62, 595–599. [Google Scholar] [CrossRef]
  337. Bhowmik, P.; Konkin, D.; Polowick, P.; Hodgins, C.L.; Subedi, M.; Xiang, D.; Yu, B.; Patterson, N.; Rajagopalan, N.; Babic, V.; et al. CRISPR/Cas9 gene editing in legume crops: Opportunities and challenges. Legume Sci. 2021, 3, e96. [Google Scholar] [CrossRef]
  338. Badhan, S.; Ball, A.S.; Mantri, N. First Report of CRISPR/Cas9 Mediated DNA-Free Editing of 4CL and RVE7 Genes in Chickpea Protoplasts. Int. J. Mol. Sci. 2021, 22, 396. [Google Scholar] [CrossRef]
  339. Razzaq, M.K.; Akhter, M.; Ahmad, R.M.; Cheema, K.L.; Hima, A.; Karikari, B.; Raza, G.; Xing, G.; Gai, J.; Khurshid, M. CRISPR-Cas9 based stress tolerance: New hope for abiotic stress tolerance in chickpea (Cicer arietinum). Mol. Biol. Rep. 2022, 49, 8977–8985. [Google Scholar] [CrossRef]
Figure 1. Diagrammatic representation of effects of drought stress on chickpea.
Figure 1. Diagrammatic representation of effects of drought stress on chickpea.
Life 12 01846 g001
Figure 2. Diagrammatic representation of scopes of CRISPR/Cas9 applications and possibilities in chickpea improvement.
Figure 2. Diagrammatic representation of scopes of CRISPR/Cas9 applications and possibilities in chickpea improvement.
Life 12 01846 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Asati, R.; Tripathi, M.K.; Tiwari, S.; Yadav, R.K.; Tripathi, N. Molecular Breeding and Drought Tolerance in Chickpea. Life 2022, 12, 1846. https://doi.org/10.3390/life12111846

AMA Style

Asati R, Tripathi MK, Tiwari S, Yadav RK, Tripathi N. Molecular Breeding and Drought Tolerance in Chickpea. Life. 2022; 12(11):1846. https://doi.org/10.3390/life12111846

Chicago/Turabian Style

Asati, Ruchi, Manoj Kumar Tripathi, Sushma Tiwari, Rakesh Kumar Yadav, and Niraj Tripathi. 2022. "Molecular Breeding and Drought Tolerance in Chickpea" Life 12, no. 11: 1846. https://doi.org/10.3390/life12111846

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