Next Article in Journal
Potential Utilization of Diluted Seawater for the Cultivation of Some Summer Vegetable Crops: Physiological and Nutritional Implications
Next Article in Special Issue
Seed Priming with Sulfhydral Thiourea Enhances the Performance of Camelina sativa L. under Heat Stress Conditions
Previous Article in Journal
Quality Characteristics of Spelt Pasta Enriched with Spent Grain
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Heat Stress in Cotton: A Review on Predicted and Unpredicted Growth-Yield Anomalies and Mitigating Breeding Strategies

1
Department of Plant Breeding and Genetics, University of Agriculture, Faisalabad 38400, Pakistan
2
Centre of Agricultural Biochemistry and Biotechnology, University of Agriculture, Faisalabad 38400, Pakistan
3
Department of Biotechnology, Chonnam National University, Chonnam 59626, Korea
4
State Key Laboratory of Cotton Biology, Institute of Cotton Research Chinese Academy of Agricultural Science, Anyang 455000, China
5
US Department of Agriculture, Agricultural Research Service, College Station, TX 77845, USA
6
School of Agriculture Sciences, Zhengzhou University, Zhengzhou 450000, China
7
Institute of Molecular Biology and Biotechnology, Bahauddin Zakariya University, Multan 60800, Pakistan
*
Authors to whom correspondence should be addressed.
Agronomy 2021, 11(9), 1825; https://doi.org/10.3390/agronomy11091825
Submission received: 4 August 2021 / Revised: 8 September 2021 / Accepted: 9 September 2021 / Published: 12 September 2021

Abstract

:
The demand for cotton fibres is increasing due to growing global population while its production is facing challenges from an unpredictable rise in temperature owing to rapidly changing climatic conditions. High temperature stress is a major stumbling block relative to agricultural production around the world. Therefore, the development of thermo-stable cotton cultivars is gaining popularity. Understanding the effects of heat stress on various stages of plant growth and development and its tolerance mechanism is a prerequisite for initiating cotton breeding programs to sustain lint yield without compromising its quality under high temperature stress conditions. Thus, cotton breeders should consider all possible options, such as developing superior cultivars through traditional breeding, utilizing molecular markers and transgenic technologies, or using genome editing techniques to obtain desired features. Therefore, this review article discusses the likely effects of heat stress on cotton plants, tolerance mechanisms, and possible breeding strategies.

1. Introduction

Upland cotton (Gossypium hirsutum) is a multipurpose cash crop. The lint of cotton is the major product of this crop and is used by the textile industry for cloth manufacturing. Cotton is cultivated on an area of about 34.1 million hectares with a production of 126.5 million bales, and it is grown in more than 35 countries [1]. India is the largest grower and producer of cotton, with production of ~6.1 million tons of cotton every year. Other leading cotton producing countries include China, United States, Brazil, Pakistan, Turkey, and Uzbekistan. China is the largest consumer of cotton, with consumption of ~7.60 million tons of cotton annually [2]
From sowing to harvesting, the cotton crop faces numerous problems including infestation of insect pests, diseases, heat, drought, cold and salinity stresses, trash during picking, and post-harvest management problems [3,4,5,6]. Each of these causes significant reduction in yield and quality of cotton fibres. Therefore, comprehensive research on each aspect is required in order to understand these problems. The present discussion focuses on high temperature stress and on minimizing the losses due to this abiotic stress. Thus, a detailed review about the effects of heat stress on cotton plants and possible strategies for its mitigation is described in the following paragraphs.
Heat stress is often called high temperature stress. It is one of the limiting factors in crop productivity. Heat stress is defined as a condition when the temperature is high enough for a sufficient period of time to cause irreversible damage to plant development and functions [7]. A sudden increase of 5–7 °C in maximum temperature for a few days with a corresponding increase in ambient minimum temperature causes “high temperature stress” in plants. The temperature requirement varies from species to species, and it also depends upon time of exposure, intensity of exposure, air or soil temperature, night/day temperature, and age of the plant. Therefore, a particular temperature cannot be defined as a cardinal point for heat stress. Generally, cool season plant species are more vulnerable to heat stress than compared to warm season crops [7,8]. Moreover, plants of similar species adapted in different climatic regions have different degrees of temperature tolerance. For example, cotton grown in the United States and China is considered to be under heat stress when temperatures increase above 38 °C, while in Pakistan and India this temperature is considered optimum and temperatures greater than 46 °C are considered as heat [9,10,11]. Cotton responses to heat stress are presented in detail below.
Traditional breeding strategies have been utilized to incorporate heat stress tolerance into upland cotton. Most of the improved cultivars and breeding lines are the outcome of purely classical breeding as very little molecular and genomic tools have been used to date. Such breeding efforts were based on intensive selection, which reduced genetic variability in cotton. Mutagenic agents have been used to create variability and have resulted in the release of high-yielding cultivars, including NIAB-78 as a successful example from Pakistan [12,13]. Due to the increased resources needed to screen large populations, reduced frequency of desirable alleles, and pleiotropic effects, mutation breeding has gradually been replaced with marker assisted breeding and site-specific mutagenesis technologies. The use of advanced genomics and biotechnological tools has also become important as the challenges to cotton production escalating. Later in this review, advanced breeding tools are discussed that can be utilized to mitigate the effect of changing climatic conditions, especially high temperature stress.

2. Effects of High Temperature Stress on Cotton

High temperatures in arid and semi-arid regions of the world are inducing negative impacts on growth, development, and productivity of several field crops [14]. Heat stress can cause damage to a cotton plant in almost every stage of its life, but it is reported that the reproductive stages of cotton are more sensitive to high temperature than compared to vegetative growth stages [15]. Both day and night temperatures play an important role in determining yield potential in crop plants, but high night temperature reduces yield and causes significant damage, while the role of high day temperature is secondary [16]. The adverse effects of high night/day temperature on different plant stages are shown in Figure 1.

2.1. Effects on Germination of Cotton Seed

Successful germination of seed and development of seedlings requires a good soil environment, especially soil moisture content and soil temperature. These requirements vary from plant species to species [17,18]. Generally, high temperature results in poor seed germination in field crops. The germination of seed is a complex physiological process that depends upon the activity of several cellular organelles and enzymes. These enzymes need to be produced in a continuous manner to perform various activities related to metabolic processes. For instance, in maize at high temperatures, these enzymes denature or slow down these activities, which results in reduced metabolism processes [19]. Moreover, high soil temperature also increases the rate of transpiration, which reduces the water availability to seed. Less availability of water also slows down the seed germination [20]. The optimal temperature of 28 to 30 °C is needed for germination of cotton seed. As the temperature decreases, the rate of seed germination also decreases, and very poor germination can be observed at temperatures <20 °C. Similarly, increases in temperature from the optimal range (≥38 °C) have also been reported to result in decreased germination of cotton seed [21].

2.2. Effects on Early Vegetative Growth Stages

In addition to germination, uniform and rapid emergence of seedlings from soil is necessary to achieve vigorous growth of plants and high yield [22]. High temperature stress during the early stages of plant growth affects the emergence of seedlings and results in the development of poor seedlings. In a previous experiment, researchers grew seedlings of cotton under various temperature regimes, i.e., 38 °C and 32 °C for 8 days. They recorded 152 mm shoot and 173 mm root growth at 32 °C temperature, while at 38 °C only 50 mm shoot and 86 mm root growths were observed [17]. The emergences and growths of cotton seedlings were also assessed for four different temperatures, i.e., 20, 30, 40, and 50 °C. The maximum emergence of seedling with vigorous growth was observed at 30 °C. The genotypes developed for hot tropical environments with higher seed weight showed vigorous seedlings even at 40 °C than compared to genotypes with smaller seed size and weight. Cotton seedlings did not emerge at 50 °C [23].
Along with the above ground parts of plants, roots are also severely affected by high temperature stress. The damaging of roots minimizes their uptake of nutrients and water from soil that can disturb the entire physiological mechanism of the plant and limit its productivity. The cotton plant has a taproot system. Studies showed that cotton roots develop poorly under high soil temperature and result in a poor crop stand [24]. A study was conducted on 10 upland cotton genotypes in order to determine the heat stress ability of roots by measuring related parameters. After 120 min of seed germinations, the seedlings were subjected to seven different temperature regimes between 25 to 45 °C. The results revealed that all the genotypes showed normal root growth up to 35 °C. The roots’ growth declined when temperature exceeded from 35 °C, and irreversible damage to roots of all the cultivars occurred at 45 °C [25].

2.3. Effects on Lateral Vegetative Growth Stages

Upland cotton is generally sown at the start of the summer season in tropical and sub-tropical parts of the world. Therefore, seedlings of cotton experience lower temperatures than compared to later lateral growth stages. The cotton crop normally experiences its highest temperature at bud initiation. This is the first flowering stage that causes instability in production [26,27]. A significant reduction in the number of fruiting branches has been observed under heat stress [28]. Dropping of floral buds, flowers, and bolls are common when the temperature increases above average. The retention of fruits on fruiting branches declined rapidly at temperatures increased to ≥33 °C in Mississippi, USA [29]. The shedding of fruiting bodies is a natural mechanism of the cotton plant for decreasing the fruit load in order to adjust the supply and demand balance of nutrients and water in the plant. Generally, the cotton plant drops ~50% premature bolls, but this percentage increases with increases in temperature. Researchers claimed that retention of bolls until harvest is the bigger challenge than compared to increasing the total number of fruiting bodies or bolls per plant. They indicated that the best possible strategy to increase cotton production is to minimize flower and boll shedding [30].
Boll weight and boll size are also reduced under heat stress conditions. Temperatures greater than 40 °C result in shortening the boll maturation period and formation of smaller sized bolls having less weight than bolls developed under normal temperatures [31]. An experiment was carried out with 23 accessions of cotton, which were grown under normal and heat stress conditions for two years. The boll weight of all the genotypes was reduced under the heat stress condition [32]. The reduction in boll weight of cotton cultivars grown under high temperatures was also observed by other researchers [33,34]. It was found that foliar application of ascorbic acid increased the heat tolerant ability of plants and minimized losses due to heat stress. Zeiher and Brown conducted several experiments under various environments, including greenhouse, growth chamber, and field environments. They concluded that boll size, fruit retention percentage, and number of seeds per boll were also reduced when temperatures were above 30 °C [29,35,36,37].

2.4. Effects on Yield and Quality of Fibre

Fibre production upon maturity is the ultimate goal of growers. Both the quantity and quality of fibre determine the end value of the cotton crop. The yield of seed cotton is a complex trait, which is the result of various morphological and physiological features of the plant. It also has a strong correlation with antioxidant activities. Both of these parameters are polygenic in nature and highly influenced by environmental conditions [38]. The negative association of fibre quantity with its quality is another challenge for breeders in terms of improving both features simultaneously [39]. Numerous studies have reported the effect of high temperature on fibre yield and quality [40,41,42,43]. An experiment was conducted on three cotton cultivars grown in nine diverse environments in order to investigate the effects of heat stress. The results revealed that lint yield and quality attributes were severely affected by high temperature. It was further concluded that cotton plants can perform well within specific temperature ranges as low temperature also adversely affected the cotton yield. The minimum threshold temperature of 22 °C was recorded to maintain seed cotton yield [44]. Another experiment revealed that both short and long term increases in day temperature decrease the biomass of cotton fruits, which ultimately results in low yield and poor quality fibres [45]. An increase of 2–3 °C temperature from the optimal temperature (32 °C) in Nanjing, China, resulted in a decrease of 10% of biomass, while yield declined by 40%. The results also revealed that the micronaire value of fibre increased under heat stress, which results in coarse fibres with less strength [46]. The developmental process of cotton fibres is sensitive to increases in temperature. Such conditions also reduce the time required for boll maturation that results in short fibres. High temperature conditions lasting up to 5 days did not change fibre quality, but the prevalence of these conditions for more than a week may cause irreversible damage to cellulose and significant reduction in fibre quality [47].

2.5. Effect on Floral Parts

Information about pollen viability and stigma receptivity is important for increasing productivity because effective pollination is essential for fruiting and seed setting in crop plants [48]. The germination and length of pollen tubes in cotton were monitored in 12 cultivars grown under different temperature regimes. The results indicated that maximum pollen germination and longest pollen tube length were observed at 32 °C. The pollen failed to germinate at 47 °C, while no pollen tubes formed above 44 °C [49]. Another experiment revealed that length of the pollen tube decreases as temperature increases above 32 °C, while germination of pollen decreases above 37 °C. Therefore, it can be suggested that the formation of the pollen tube is more sensitive to high temperature stress than compared to pollen germination. G. barbadense was found to be more sensitive to heat stress. A lower in vitro pollen germination percentage was recorded for heat treated pollen of pima cotton than compared to upland cotton [50]. The length of the filament was significantly reduced when cotton flowers were exposed to high temperature stress, and this resulted in the appearance of an elongated stigma. The actual length of stigma remained the same, but the length of filaments were reduced to the extent that the stigma appeared to be very long, and the process of self-pollination was badly affected [29]. The loss of receptivity of the stigma under high temperature is also reported in sweet cherry and peach [51,52]. The penetration of pollen grains to the ovule via the pollen tube ensures the successful fertilization process, and the penetration of pollen tube through the stigma, style, and ovule at high temperatures was reduced, which resulted in poor seed setting [53]. The effects of heat stress on various reproductive phases of plants are illustrated in Figure 2.

2.6. Effects on Physiological Attributes

Plant growth and development is based on physiological processes such as photosynthesis, membrane permeability, and stomatal conductance [54]. Heat stress adversely affects physiological attributes of plants and limits productivity [55]. The rate of photosynthesis is significantly reduced or even inhibited at high temperatures. The deactivation of ribulose 1,5 bisphosphate carboxylase/oxygenase (Rubisco) and increase in ionic conductance of thylakoid membranes are the primary causes of photosynthesis reduction or inhibition in cotton during high temperature conditions [56,57]. Chlorophyll content also decreases when cotton plants are exposed to high temperature that results in a decrease in the rate of photosynthesis [58]. Heat stress changes the permeability of membranes and alters cell differentiation and elongation by causing injuries to cellular membranes and deforming the organization of microtubules and cytoskeleton [59]. Higher cell membrane thermostability is positively associated with heat tolerance in cotton. Therefore, a number of experiments have been conducted to screen for heat tolerant accessions on the basis of cell membrane thermostability [60,61,62]. Stomatal conductance is directly related to water relations and photosynthesis in plants. High temperature causes opening of stomata and an increase in stomatal conductance that results in a decrease in the water potential of leaf. High stomatal conductance also increases the rate of transpiration and intercellular CO2. Stomatal conductance increases up to 40% in most of plant species when temperature rises from 30 to 40 °C [63]. The advantage of higher stomatal conductance is associated with cooling of leaves, which provides tolerance to heat stress [64]. Experimental results have shown that upland cotton has more stomatal conductance and higher rate of photosynthesis under high temperature conditions than compared to pima cotton [65]. Studies revealed that differences in cotton accessions and species for stomatal conductance are under genetic control. Thus, this trait can be improved through breeding and selection [66,67,68].

3. Mechanisms of Heat Tolerance

3.1. Antioxidant Activity in Response to Oxidative Stress

Crop plants face environmental stresses which alter the various metabolic activities in order to ensure balance between production and consumption of energy through oxidation and reduction reactions [69]. This change in metabolism alters the concentration of various molecules. Likewise, metabolic imbalances during high temperature stress promote the extra-accumulation of reactive oxygen species (ROS) into the cellular compartment of plants [70]. ROSs are highly reactive chemicals formed from O2. Excessive production and over-accumulation of ROS in plant cells cause irreversible damage to its organelles through oxidative stress. However, a balanced amount of ROS is required for normal activities such as detoxification of poisonous substances, antimicrobial phagocytosis, and apoptosis. ROS also benefits plants by acting as signaling molecules for activation of numerous genes related to stress tolerance, cell proliferation, seed germination, growth of root hairs, and cell senescence [71,72]. The over-accumulation of ROS during stress conditions results in the oxidation damage of vital molecules such as DNA, proteins, and lipids. This condition is termed as oxidative stress in plants [73]. ROS includes free radicals such as the hydroxyl radical (OH) and superoxide anion (O2), as well as non-radicals such as singlet oxygen (1O2) and hydrogen peroxide (H2O2). These species are produced by excitation and reduction in intra-cellular oxygen (O2). The schematic diagram of ROS production is illustrated in Figure 3.
The concentration of ROS and scavenging capacity of antioxidants in cotton is considered as a selection criterion for heat tolerant accessions [74]. An experiment was conducted using two cotton cultivars, and heat stress was applied gradually from 30 to 45 °C on 30 days old seedlings. The results revealed a 206 to 248% increase in hydrogen peroxide content and a 40 to 170% increase in lipid peroxidation under heat stress conditions. The concentration of non-enzymatic antioxidants also increases with increases in temperature, while the activity of enzymatic antioxidants such as superoxide dismutase (SOD), catalase (CAT), peroxidase (POX), and ascorbate peroxidase (APX) increase 56–70%, 37–69%, 43–91%, and 22–27%, respectively. It was concluded that genotypic differences exist in cultivars for ROS production and antioxidants response. Higher levels of antioxidants and lower levels of ROS during high temperature are an indication of heat stress tolerance [75]. In another study, the cotton plants were grown at two temperature regimes, i.e., 38 and 45 °C. The results indicated non-significant differences in the concentration of hydrogen peroxide at both temperatures, while the concentration of proline decreased rapidly and significantly as the temperature increased from 30 to 45 °C. The activity of SOD declined at 45 °C while the activity of CAT, POX, and APX increased with the increase in temperature [76]. It is found that the exogenous application of hydrogen peroxide on cotton plants triggers the activity of SOD and CAT. It was further concluded that foliar applications of H2O2 on field grown cotton can enhance the heat tolerant ability without compromising yield [77]. The effect of high night temperature on biochemistry of leaf and pistil was studied in upland cotton cultivar. The results indicated that glutathione reductase activity in leaves is increased with an increase in night temperature, while no change in the concentration of glutathione reductase was observed in pistils. This shows that the antioxidant mechanism of pistil or floral parts is less sensitive to changes in mean night temperature than compared to leaves or vegetative parts of the cotton plant [78].

3.2. Role of Heat Shock Proteins

Heat shock proteins (HSPs) are important contributors to cellular homeostasis under heat stress. These molecular chaperons are upregulated during high temperature conditions and perform various activities in order to maintain the integrity of the cell [79]. On the basis of molecular weight, HSPs are divided into five major groups, i.e., small HSPs, HSP60, HSP70, HSP90, and HSP100 [80]. The details of each group are summarized in Table 1. Each HSP group is unique in nature and specific in function. Small HSPs (sHSPs) have low molecular weight (12–40 kDa) and are the most diverse in nature with respect to cellular location, function, and sequence similarities [81]. These sHSPs bind to non-native proteins, prevent non-native aggregation through hydrophobic interactions, and facilitate their refolding by ATP-dependent chaperons such as ClpB/DnaK [82]. Almost all sHSPs have an α-crystallin domain, which forms a dodecamer double ring and helps in the folding of proteins [83]. Previous work has revealed that the expression of the sHSP coding gene, i.e., Hsp 17.7, is directly related to thermal stress tolerance in plants [84]. A quantitative expression analysis of GHSP26 (a small HSP coding gene in cotton) indicated that the leaves of cotton have 100-fold increased concentration of proteins encoded by this gene during water deficit conditions [85].
HSP60 is generally known as a mitochondrial chaperon or chaperonin 60. It plays two essential roles in mitochondria during high temperature conditions, i.e., maintenance of the unfolded state of proteins for their transportation across the inner mitochondrial membrane and the folding of important proteins into a matrix [88]. HSP60 is also involved in assisting proteins that help in photosynthesis such as Rubisco [99]. Studies revealed that a mutation in Chaperonin-60α gene that codes for HSP60 protein causes a defection in chloroplasts, which ultimately results in poor seedlings and embryo development in Arabidopsis plant [100]. However, deletion of this gene results in cell death [101]. It has been experimentally verified that transgenic tobacco plants with reduced Cpn60β (chaperonin 60β) exhibited phenotypic defects such as delayed flowering, stunted growth, and leaf chlorosis [102]. HSP70s are considered important cellular machinery involved in the folding of proteins and in preventing their aggregation [103]. The overexpression of HSP70s is an indication of heat tolerance in plants. It is reported that HSP70 genes of cotton play essential roles during fibre development. The inhibition of these genes results in the retardation of fibre elongation. The inhibition of HSP70 genes results in oxidative stress by elevating the level of H2O2, which causes damage to epidermal layers of the ovule [104]. HSP70 proteins also act as signaling molecules for transcriptional activation and de-activation [91].
HSP90 proteins are quite distinct from other chaperons because most of them are substrates involved in signal transduction, such as signaling kinases and hormone receptors [105]. They also manage the folding of proteins [106]. HSP90s are among the most abundant proteins of the cell (1–2% of the total), are constitutive in nature, and act together with HSP70s as multi-chaperone machinery. The expression of HSP90 proteins increases significantly during hot conditions [107]. HSP100 belongs to the AAA ATPase family and performs various functions such as unfolding and disaggregation of proteins [108]. In addition to heat stress tolerance, HSP100 also performs housekeeping functions in the cell, including the development of chloroplasts [109,110].

3.3. Small RNAs Activity in Regulating Heat Stress

With the advancement in high throughput sequencing technologies, plant researchers have revealed various roles of microRNA (miRNA) and small interfering RNA (siRNA) during biotic and abiotic stress conditions. These small RNAs are actively involved in the degradation of mRNA and prevent translation of various proteins in plant cells [111,112,113]. The regulatory roles of various miRNAs have been characterized during high temperature conditions in many plant species, including chestnuts and Arabidopsis, e.g., the miR156 and miR157 families comprise 17 and four miRNAs, respectively. These miRNAs upregulate during hot temperature conditions and target SPL genes, which are essential for floral development in plants. Thus, flowering is controlled by these miRNAs when plants are exposed to high temperature conditions [114,115]. Over-production of miR157 targets SPL genes in cotton during heat stress results in a reduction in flowers and smaller sized bolls with fewer seeds [116]. The similar role of these miRNA families is reported in Brassica rapa [117], citrus [118], and Arabidopsis [119].
Experiments revealed that the expression of miR159 is down-regulated in heat tolerant genotypes of wheat upon exposure to hot temperature regimes. This miRNA acts as a negative regulator of MYB transcriptional factors [120]. The major role of the auxin response factors (ARFs) gene family is the regulation of auxin levels in plants. The over-expression of miR160 in cotton increases its susceptibility to high temperature stress by suppressing the expression of ARF genes [121]. It is found that the expression of miR162 increases up to 15-fold during drought conditions. Increases in concentration of miR162 under heat and salinity stresses is reported in cotton and rice [122,123]. This microRNA controls the transcription of numerous genes by targeting zinc finger proteins (ZFPs) and acts as a regulator of dicer such as proteins [124,125]. The expression level of miR164 was observed to decrease 0.3-fold under heat stress conditions. It targets HSP17 genes and also regulates the expression of various genes essential for mitogen activated protein kinase (MAPK) mediated signaling pathways and the activation of NAC transcriptional factors in wheat, rice, and alfalfa [126,127,128]. It is reported that expression of miR171 increased several folds upon exposure of the plant to high temperature. It targets GRAS genes, which are involved in various developmental processes, i.e., flowering time, floral meristem determinacy, plant height, and leaf architecture in cotton [129,130].
The regulation of flowering and floral organs is controlled by AP2 genes [131]. This gene family is regulated by miR172, as reported in roses [132]. Thus, up-regulation and downregulation of these miRNAs is directly related to flowering timing and the transition of floral and vegetative phases in rice and Arabidopsis during high temperature conditions [133,134]. The miR390 of cotton controls the formation of lateral roots by targeting ARF genes [135]. The miR393 is considered a regulator of auxin receptors [136]. Over-expression causes a delay in flowering and results in poor development of roots. Thus, decreased levels of miR393 are an indicator of stress tolerance in cotton and rice [137,138]. F-box proteins perform various activities during stress conditions such as degradation of proteins, rolling, and senescence of leaves [139,140]. It was found that miR394 prevents the translation of F-box genes family transcripts during abiotic stresses to maintain the optimal levels of proteins for normal functioning of plants, particularly in rice and Arabidopsis [141,142]. The miR395 regulates APS genes in response to various abiotic stresses. The major function of this gene family is the assimilation of sulphate [143].

4. Breeding Strategies for High Temperature Stress Tolerance

4.1. Conventional Breeding Approaches

Assessment of germplasm is a prerequisite for breeding stress tolerance. Numerous experiments have been conducted to identify heat stress tolerant genotypes from the available genepool. Moreover, the utilization of crop wild relatives is also gaining popularity in plant breeding due to their novel features that are lacking in domesticated cultivars. Most of these novel features are related to biotic and abiotic environmental stress. It is recommended to screen related wild species and relatives in order to have a diverse gene pool [144]. Although gene transfer from wild to cultivated species encounters numerous problems and is not always possible without recombinant DNA technology, the rapidly evolving technologies in plant sciences have made it quite possible to transfer genes among many species, as is discussed below [145]. After the identification of a suitable gene and trait, the next step is to transfer it to a desirable genotype or to purify the identified plant through selection. For this purpose, single plant selection, bulk selection, and pedigree selection are among the most widely used classical breeding methods in cotton [146,147]. These methods are used in cotton improvement along with molecular breeding tools for quick and efficient screening and genetic gain.

4.2. Molecular and Biotechnological Approaches

In addition to conventional screening and breeding approaches, molecular markers and biotechnological tools are also useful for improving stress tolerance of cotton genotypes [148]. Numerous markers such as amplified fragment length polymorphism (AFLP) and randomly amplified polymorphic DNA (RAPD) markers have been successfully utilized screening cotton genotypes for heat tolerance in the past [149,150]. Currently, simple sequence repeats (SSRs) and single nucleotide polymorphism (SNPs) are widely used markers for identifying quantitative traits loci (QTLs) related to stress tolerance in cotton [151,152]. The experiments were conducted by using heat tolerant and susceptible cultivars to determine heat responsive genes in upland cotton. Twenty-five expressed sequence tags (ESTs) were sequenced to study the homology of genes. The expression level of a few ESTs was also quantified using real time PCR. The results indicated that expression of folylpolyglutamate synthase (FPGS3) and IAA-ala hydrolase (IAR3) coding genes was significantly up-regulated during long-term and short-term high temperature stress. The expression of two non-annotated ESTs, i.e., GhHS128 and GhHS126, was also found to be up-regulated under hot conditions. Thus, it was suggested that these two non-annotated ESTs are heat tolerant candidate genes [153].
In order to investigate the molecular mechanism of high temperature stress tolerance, the expression of some heat responsive genes was quantified through real time PCR in tolerant and susceptible upland cotton cultivars. The genes belong to various groups, i.e., HSFA1b and HSFA2 are heat stress transcriptional factors; HSP101, HSP70-1, and GHSP26 code for heat shock proteins; ANNAT8 is involved in calcium signaling; and APX1 controls antioxidant activity. The level of GHSP26 increased in all genotypes, while the expression of HSP101 and HSP70-1 increased several-fold only in the seedlings of heat tolerant cultivars. The expression of APX1 increased significantly in a heat-tolerant cultivar (VH-260), indicating the involvement of antioxidant activity in conferring heat tolerance. No significant change in the expression of ANNAT8 was observed in heat susceptible cultivars. The expression of HSFA2 and HSFA1b was several folds higher in leaves and ovaries of heat tolerant accessions than compared to heat susceptible accessions [154]. In order determine the SNP markers linked to the mitochondrial small heat shock protein (MTsHSP), a study was conducted by using accessions belonging to various cotton species, i.e., G. aridum, G. sturtianum, G. gossypioides, G. stocksii, G. arboreum, G. laxum, and G. herbaceum. Approximately 21 SNPs were identified for this gene by using PCR cloning and sequencing techniques, which could be useful for cotton improvement [155]. Transcriptomic analysis of 82 genes belonging to the GhHSP20 family revealed their involvement in developmental and physiological processes of cotton. Most of them were regulatory in nature and expressed only under hot conditions, while eight genes were found to be involved in conferring tolerance for multiple stresses, namely heat, drought, and salinity [156].
Rapid advancements in applied genomics have resulted in useful tools for plant improvement. For example, markers linked to known genes or QTLs can be used for marker-assisted selection (MAS), as well as for genomic selection. Genomic selection assists the breeders in utilizing the molecular markers in the absence of phenotypic data. It can reduce the time for cultivar development through more efficient selection of progeny in early generations. An experiment was conducted with 550 recombinant inbred lines of cotton, and six fiber quality traits were evaluated using genomic selection. A total of 6292 markers were obtained through genotyping by sequencing. It was revealed that genomic selection could potentially predict genomic estimated breeding value in upland cotton fiber quality attributes [157]. Association mapping is also an effective technique for cotton improvement when information on population structure and linkage disequilibrium (LD) is available. This method is quite useful for reducing the laborious work involved in screening large populations [158]. Genome-wide association studies (GWAS) represent a powerful approach for identifying the locations of genetic factors that underlie complex traits. GWAS has been successfully implemented in cotton for the identification of single nucleotide polymorphism (SNP) loci and candidate genes for various attributes [159].

4.3. Transgenic Approaches

Transgenic techniques have also been extensively used to improve cotton cultivars for better tolerance with respect to high temperature stress. Recently, heat shock protein 70 (AsHSP70) from Agave sisalana was transformed in cotton through the Agrobacterium mediated transformation method. Expression studies showed the higher expression of transformed gene in different plant tissues under high temperature. Additionally, transgenic cotton plants exhibited improved performance for the measurable physiological and biochemical indicators [160]. In another study, over-expression of both AVP1 and OsSIZ1 genes in cotton improves lint yield compared to wild-type cotton under combined drought and heat stress conditions, and it does not have any negative affect on overall cotton yield when there is no stress. Furthermore, transgenic cotton plants also had 72% more photosynthetic rates two hours before the outset of heat stress and 108% higher photosynthetic rates during heat stress [161]. The function of Arabidopsis heat shock protein 101 (AtHSP101) is well known in heat tolerance at a vegetative stage, and the overexpression of this protein in cotton (Coker-312) clearly exhibited the increased germination percentage and enhanced pollen tube elongation under high temperatures than compared to non-transgenic cotton [162]. Therefore, improved heat tolerance of reproductive systems in transgenic cotton is crucial for enhanced yield on a sustainable basis in the face of climate change. Conversely, the Arabidopsis SUMO E3 ligase (AtSIZ1) is a key gene for plant heat stress response, as AtSIZ1 mutant plants exhibited increased susceptibility to high temperatures. In cotton, the overexpression of OsSIZ1 gene, a rice homolog of AtSIZ1, conferred tolerance to both drought and heat stresses than compared with non-transgenic plants by demonstrating enhanced net photosynthesis rate and improved growth and development [163]. In another study, research revealed that the ectopic expression of Arabidopsis stress associated gene (AtSAP5) in transgenic cotton (Coker-312) protects several components of carbon gain and growth under extreme drought and associated heat stress conditions [164].
However, in order to render transgenic cotton more acceptable and efficient, it is necessary to focus on improving transformation efficiency. In a nutshell, these studies indicate that integrating heat stress-related genes in cotton is a viable strategy for engineering high-temperature tolerant cotton cultivars that could significantly improve lint yields in marginal environments, resulting in a sustainable cotton production. The effectiveness of heat stress responsive genes, for which its expression could be successful without the accompanying yield penalties, would determine the end degree of success.

4.4. CRISPR-Cas Mediated Genome Editing

Tolerance relative to high temperatures is mostly regulated by many genes based on the degree of stress and the stress tolerance mechanism. It will be more difficult to target a tolerance mechanism that is controlled by multiple genes. Plant heat stress response is precisely regulated by a complex web of TFs from distinct families. These TFs improve plant heat stress tolerance by modulating the expression of several stress responsive genes, either individually or in conjunction with other regulatory factors. There are numerous successful genetic engineering applications inducing heat stress tolerance in plants using heat stress TFs and HSPs genes [165,166]. However, genetically modified crop plants are subject to stringent regulatory requirements, which may cause lab research to be delayed in reaching the market. As an alternative to traditional transgenic approaches, recently emerged CRISPR-Cas-mediated genome editing allows researchers to alter, modify or swap alleles, and insert or silence gene(s) in a predefined manner [167].
High temperatures alter the expression pattern of several plant genes either by upregulating or down-regulating them. Although our understanding of differentially expressed genes in response to drought and salt stress has expanded, relatively less focus has been made on studying heat stress associated genes in cotton. Studying the expression pattern of heat stress responsive genes in cotton under long-term heat stress clearly showed that expressions of HS126, HS128, FPGS, TH1, and IAR3 genes increased under high temperature. In contrast, the expressions of ABCC3, CIPK, CTL2, LSm8, and RPS14 genes were down-regulated [153]. Therefore, targeted modulation of these up-regulated and down-regulated genes in cotton using the CRISPR-Cas system would be an exciting opportunity for combating the negative impact of heat stress. Moreover, multiple HSPs and TFs associated with heat stress sensitive genes have been proposed as potential candidates for improving plant heat tolerance [168]. Therefore, understanding the exact role of these genetic regulators paves the way for the development of enhanced heat tolerance, while maintaining overall plant resilience. In maize plants, peak photosynthesis has been observed between 20 and 32 °C, and the subsequent increase in temperature caused a decrease in net photosynthesis rate depending on plant growth stage [169]. CRISPR/Cas9 mediated disruption of the heat-stress sensitive albino-1 (HSA1) gene in rice exhibits higher heat sensitivity compared to wild plants [170]. The slagamous-like 6 (Slagl6) gene was reported as a potential gene in the study of facultative parthenocarpy. Thus, researchers have successfully developed heat-tolerant parthenocarpic tomato fruit by mutating the SlAGL6 gene using the state-of-the-art CRISPR-Cas9 system [171]. In addition, thermosensitive male sterile maize lines have also been developed by mutating the thermosensitive genic male-sterile 5 (TMS5) gene using CRISPR-Cas9 editing [172].
CRISPR-Cas9 has been modified and exploited for a range of new functions, including controlling gene regulation by activating and suppressing target gene expression using CRISPR activation and interference systems [173]. Positive gene regulators associated with HSPs and stress related TFs could be activated through the CRISPR activation system with high specificity. Moreover, negative regulators could be knocked out by using the CRISPR interference system. In one of study, the BZR1 gene was up-regulated and repressed using CRISPR activator and interference systems. The results show that the overexpression of the BZR1 gene enhances H2O2 production and recovery of thermo tolerance in rice, while plants with suppression of the gene show impaired production of H2O2 in apoplast and reduced heat tolerance [174]. Previously, the roles of MAP3Ks remained poorly understood in cotton. Recently, it has been reported that MAP3K65 gene expression is induced by multiple signaling molecules, pathogen infection, and heat stress. This gene enhances susceptibility to pathogen infection and heat stress by negatively modulating growth and development related processes. Moreover, silencing GhMAP3K65 enhanced resistance to pathogen infection and heat stress in cotton. Therefore, GhMAP3K65 is a potential candidate gene to target with the CRISPR-Cas9 genome editing system in order to engineer heat tolerance in cotton [175]. In conclusion, maneuvering positive and negative regulators of heat stress signaling molecules in cotton could, thus, be exploited to develop new cotton cultivars tolerant to extreme temperatures.

5. Conclusions and Future Directions

High temperature exerts a negative impact on cotton growth and yield in all stages of plant growth. It reduces lint quantity and quality by impeding normal plant biological processes and pathways. Understanding the heat tolerance mechanism and molecular characterization of related genes is essential for developing stress tolerant cultivars for sustainable cotton production under changing climatic conditions. Undoubtedly, the heat stress TFs, HSPs, and other genes are important for maintaining the secondary structure of proteins, and rapid sensing of heat is also critical to induce the protective mechanisms against high temperature stress. Traditional breeding approaches for developing stress tolerance are being complemented by new technologies, including state-of-the-art genome editing tools, speed breeding approaches, and various omics tools, which may decrease the time needed to develop cotton cultivars with increased heat tolerance. For engineering high temperature stress tolerance, the CRISPR-Cas system is considered a non-genetically modified (nGM) approach, thus allowing for the expansion of scientific community efforts to introduce heat stress tolerance in future cotton cultivars for all cotton growing regions. Furthermore, the increasing demand of high-quality lint yield in a rapidly growing world population will be confronted through concept of speed breeding. It will aid quick generation advancement in order to shorten the overall growth cycle and accelerate cotton breeding programs. In addition, recent advances in sequencing technologies have made significant strides in successfully sequencing complex cotton genome. Subsequently, applications of various omics approaches have pointedly enhanced our understanding of cotton physiology and the functions of genes in response to heat stress. Therefore, differently expressed genes, proteins, and metabolites identified through different omics tools can be used as a potential biomarker to develop high temperature-resilient cotton cultivars.

Author Contributions

Conceptualization, M.T.A., S.M., L.H., Y.J. and M.S.M.; writing and draft preparation, S.M., M.T.A. and M.S.M. writing, review and editing, M.T.A., R.M.A., S.-H.Y. and G.C.; supervision, X.D., I.A.R. and M.T.A. 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.

Acknowledgments

The presented study is part of research proposal “Genotyping and development of heat stress tolerant cot-ton germplasm having enhanced quality traits” No. 964, and the authors are grateful to the Center for Advanced Studies and Punjab Agricultural Research Board (CAS-PARB), Pakistan, for providing funds for this study.

Conflicts of Interest

The authors declare no conflict of interest.

Disclaimer

USDA is an equal opportunity provider and employer.

References

  1. Meyer, L.A. The World and US Cotton Outlook for 2019/20. 2019. Available online: https://www.usda.gov/sites/default/files/documents/Leslie_Meyer.pdf (accessed on 12 May 2021).
  2. Khan, M.A.; Wahid, A.; Ahmad, M.; Tahir, M.T.; Ahmed, M.; Ahmad, S.; Hasanuzzaman, M. World cotton production and consumption: An overview. Cotton Prod. Uses 2020, 1–7. [Google Scholar] [CrossRef]
  3. Zahra, N.; Shaukat, K.; Hafeez, M.B.; Raza, A.; Hussain, S.; Chaudhary, M.T.; Wahid, A. Physiological and molecular responses to high chilling and freezing temperature in plant growth and production: Consequences and mitigation possibilities. In Harsh Environment and Plant Resilience: Molecular and Functional Aspects; Springer: New York, NY, USA, 2021; p. 235. [Google Scholar]
  4. Zhu, Y.N.; Shi, D.Q.; Ruan, M.B.; Zhang, L.L.; Meng, Z.H.; Liu, J.; Yang, W.C. Transcriptome analysis reveals crosstalk of responsive genes to multiple abiotic stresses in cotton (Gossypium hirsutum L.). PLoS ONE 2013, 8, e80218. [Google Scholar]
  5. Downes, S.; Walsh, T.; Tay, W.T. Bt resistance in Australian insect pest species. Curr. Opin. Insect Sci. 2016, 15, 78–83. [Google Scholar] [CrossRef]
  6. Van der Sluijs, M.J.; Hunter, L. A review on the formation, causes, measurement, implications and reduction of neps during cotton processing. Text. Prog. 2016, 48, 221–323. [Google Scholar] [CrossRef]
  7. Hall, A.E.; Botany and Plant Sciences Department University of California, Riverside. Heat Stress and Its Impact. 2001. Available online: https://plantstress.com/heat/ (accessed on 13 June 2021).
  8. Dhyani, K.; Ansari, M.W.; Rao, Y.R.; Verma, R.S.; Shukla, A.; Tuteja, N. Comparative physiological response of wheat genotypes under terminal heat stress. Plant Signal. Behav. 2013, 8, e24564. [Google Scholar] [CrossRef] [PubMed]
  9. Phillips, J.B. Cotton Response to High Temperature Stress During Reproductive Development. 2012. Available online: https://scholarworks.uark.edu/etd/394 (accessed on 22 July 2021).
  10. Raza, A.; Ahmad, M. Analysing the Impact of Climate Change on Cotton Productivity in Punjab and Sindh, Pakistan. 2015. Available online: https://mpra.ub.uni-muenchen.de/72867/ (accessed on 7 July 2021).
  11. Zahid, K.R.; Ali, F.; Shah, F.; Younas, M.; Shah, T.; Shahwar, D.; Hassan, W.; Ahmad, Z.; Qi, C.; Lu, Y. Response and tolerance mechanism of cotton Gossypium hirsutum L. to elevated temperature stress: A review. Front. Plant Sci. 2016, 7, 937. [Google Scholar] [CrossRef] [Green Version]
  12. Shamsuzzaman, K.; Hamid, M.; Azad, M.; Hussain, M.; Majid, M. Varietal improvement of cotton (Gossypium hirsutum) through mutation breeding. In Improvement of New and Traditional Industrial Crops by Induced Mutations and Related Biotechnology; International Atomic Energy Agency: Vienna, Austria, 2003; pp. 81–94. [Google Scholar]
  13. Ahloowalia, B.; Maluszynski, M.; Nichterlein, K. Global impact of mutation-derived varieties. Euphytica 2004, 135, 187–204. [Google Scholar] [CrossRef]
  14. Challinor, A.; Wheeler, T.; Craufurd, P.; Slingo, J. Simulation of the impact of high temperature stress on annual crop yields. Agric. For. Meteorol. 2005, 135, 180–189. [Google Scholar] [CrossRef] [Green Version]
  15. Snider, J.L.; Oosterhuis, D.M.; Skulman, B.W.; Kawakami, E.M. Heat stress-induced limitations to reproductive success in Gossypium hirsutum. Physiol. Plant. 2009, 137, 125–138. [Google Scholar] [CrossRef]
  16. Khan, A.H.; Min, L.; Ma, Y.; Wu, Y.; Ding, Y.; Li, Y.; Xie, S.; Ullah, A.; Shaban, M.; Manghwar, H. High day and night temperatures distinctively disrupt fatty acid and jasmonic acid metabolism, inducing male sterility in cotton. J. Exp. Bot. 2020, 71, 6128–6141. [Google Scholar] [CrossRef]
  17. Nabi, G.; Mullins, C. Soil temperature dependent growth of cotton seedlings before emergence. Pedosphere 2008, 18, 54–59. [Google Scholar] [CrossRef]
  18. Zia, S.; Khan, M.A. Effect of light, salinity, and temperature on seed germination of Limonium stocksii. Can. J. Bot. 2004, 82, 151–157. [Google Scholar] [CrossRef] [Green Version]
  19. Riley, G.J. Effects of high temperature on the germination of maize (Zea mays L.). Planta 1981, 151, 68–74. [Google Scholar] [CrossRef] [PubMed]
  20. Burke, J.J.; Upchurch, D.R. Leaf temperature and transpirational control in cotton. Environ. Exp. Bot. 1989, 29, 487–492. [Google Scholar] [CrossRef]
  21. Krzyzanowski, F.C.; Delouche, J.C. Germination of cotton seed in relation to temperature. Rev. Bras. Sementes 2011, 33, 543–548. [Google Scholar] [CrossRef] [Green Version]
  22. Parera, C.A.; Cantliffe, D.J. Presowing seed priming. Hortic. Rev. 1994, 16, 109–141. [Google Scholar]
  23. Raphael, J.; Gazola, B.; Nunes, J.G.; Macedo, G.C.; Rosolem, C.A. Cotton germination and emergence under high diurnal temperatures. Crop Sci. 2017, 57, 2761–2769. [Google Scholar] [CrossRef]
  24. Carmo-Silva, A.E.; Gore, M.A.; Andrade-Sanchez, P.; French, A.N.; Hunsaker, D.J.; Salvucci, M.E. Decreased CO2 availability and inactivation of Rubisco limit photosynthesis in cotton plants under heat and drought stress in the field. Environ. Exp. Bot. 2012, 83, 1–11. [Google Scholar] [CrossRef]
  25. Sethar, M.A.; Laidman, D.; Pahoja, V.M.; Chachar, Q.; Mirjat, M.A. Thermotolerance in Cotton Roots at Early Stage of Emergence. Pak. J. Biol. Sci. 2001, 4, 361–364. [Google Scholar] [CrossRef]
  26. Reddy, K.R.; Hodges, H.F.; McKinion, J.M. A comparison of scenarios for the effect of global climate change on cotton growth and yield. Funct. Plant Biol. 1997, 24, 707–713. [Google Scholar] [CrossRef]
  27. Kamal, M.; Saleem, M.; Shahid, M.; Awais, M.; Khan, H.; Ahmed, K. Ascorbic acid triggered physiochemical transformations at different phenological stages of heat-stressed Bt cotton. J. Agron. Crop Sci. 2017, 203, 323–331. [Google Scholar] [CrossRef]
  28. Saleem, M.F.; Kamal, M.A.; Anjum, S.A.; Shahid, M.; Raza, M.A.S.; Awais, M. Improving the performance of Bt-cotton under heat stress by foliar application of selenium. J. Plant Nutr. 2018, 41, 1711–1723. [Google Scholar] [CrossRef]
  29. Brown, P. Cotton Heat Stress. 2008. Available online: https://cals.arizona.edu/azmet/az1448.pdf (accessed on 14 July 2021).
  30. Tariq, M.; Yasmeen, A.; Ahmad, S.; Hussain, N.; Afzal, M.N.; Hasanuzzaman, M. Shedding of fruiting structures in cotton: Factors, compensation and prevention. Trop. Subtrop. Agroecosystems 2017, 20, 251–262. [Google Scholar]
  31. Karademir, E.; Karademir, Ç.; Ekinci, R.; Başbağ, S.; Başal, H. Screening cotton varieties (Gossypium hirsutum L.) for heat tolerance under field conditions. Afr. J. Agric. Res. 2012, 7, 6335–6342. [Google Scholar]
  32. Rauf, S.; Khan, T.M.; Naveed, A.; Munir, H. Modified path to high lint yield in upland cotton (Gossypium hirsutum L.) under two temperature regimes. Turk. J. Biol. 2007, 31, 119–126. [Google Scholar]
  33. Kamal, M.; Saleem, M.; Wahid, M.; Shakeel, A. Effects of ascorbic acid on membrane stability and yield of heatstressed BT cotton. J. Anim. Plant Sci. 2017, 27, 192–199. [Google Scholar]
  34. Singh, K.; Wijewardana, C.; Gajanayake, B.; Lokhande, S.; Wallace, T.; Jones, D.; Reddy, K.R. Genotypic variability among cotton cultivars for heat and drought tolerance using reproductive and physiological traits. Euphytica 2018, 214, 57. [Google Scholar] [CrossRef]
  35. Zeiher, C.A.; Brown, P.W.; Silvertooth, J.C.; Matumba, N.; Mitton, N. The Effect of Night Temperature on Cotton Reproductive Development. 1994. Available online: https://repository.arizona.edu/bitstream/handle/10150/209598/370096-089-096.pdf?sequence=1 (accessed on 27 June 2021).
  36. Zeiher, C.; Matumba, N.; Brown, P.; Silvertooth, J. Response of upland cotton to elevated night temperatures. II. Results of controlled environmental studies. In Proceedings of the Beltwide Cotton Conferences, San Antonio, TX, USA, 4–7 January 1995. [Google Scholar]
  37. Brown, P.; Zeiher, C. A model to estimate cotton canopy temperature in the desert southwest. In Proceedings of the Beltwide Cotton Conferences, San Diego, CA, USA, 5–9 January 1998; National Cotton Council of America: Memphis, TN, USA, 1998. [Google Scholar]
  38. Sabagh, A.E.; Hossain, A.; Islam, M.S.; Barutcular, C.; Ratnasekera, D.; Gormus, O.; Amanet, K.; Mubeen, M.; Nasim, W.; Fahad, S. Drought and heat stress in cotton (Gossypium hirsutum L.): Consequences and their possible mitigation strategies. In Agroomic Crops; Springer: New York, NY, USA, 2020. [Google Scholar] [CrossRef]
  39. Yaqoob, M.; Fiaz, S.; Ijaz, B. Correlation analysis for yield and fiber quality traits in upland cotton. Commun. Plant Sc. 2016, 6, 55–60. [Google Scholar]
  40. Peng, S.; Krieg, D.; Hicks, S. Cotton lint yield response to accumulated heat units and soil water supply. Field Crops Res. 1989, 19, 253–262. [Google Scholar] [CrossRef]
  41. Reddy, K.R.; Davidonis, G.H.; Johnson, A.S.; Vinyard, B.T. Temperature regime and carbon dioxide enrichment alter cotton boll development and fiber properties. Agron. J. 1999, 91, 851–858. [Google Scholar] [CrossRef]
  42. Pettigrew, W. The effect of higher temperatures on cotton lint yield production and fiber quality. Crop Sci. 2008, 48, 278–285. [Google Scholar] [CrossRef] [Green Version]
  43. Abro, S.; Rajput, M.T.; Khan, M.A.; Sial, M.A.; Tahir, S.S. Screening of cotton (Gossypium hirsutum L.) genotypes for heat tolerance. Pak. J. Bot 2015, 47, 2085–2091. [Google Scholar]
  44. Liakatas, A.; Roussopoulos, D.; Whittington, W. Controlled-temperature effects on cotton yield and fibre properties. J. Agric. Sci. 1998, 130, 463–471. [Google Scholar] [CrossRef] [Green Version]
  45. Li, X.; Shi, W.; Broughton, K.; Smith, R.; Sharwood, R.; Payton, P.; Bange, M.; Tissue, D.T. Impacts of growth temperature, water deficit and heatwaves on carbon assimilation and growth of cotton plants (Gossypium hirsutum L.). Environ. Exp. Bot. 2020, 179, 104204. [Google Scholar] [CrossRef]
  46. He, X.-Y.; Zhou, Z.-G.; Dai, Y.-J.; Qiang, Z.-Y.; Chen, B.-L.; Wang, Y.-H. Effect of increased temperature in boll period on fiber yield and quality of cotton and its physiological mechanism. Yingyong Shengtai Xuebao 2013, 24, 12. [Google Scholar]
  47. Bo, X.; Zhou, Z.-G.; Guo, L.-T.; Xu, W.-Z.; Zhao, W.-Q.; Chen, B.-L.; Meng, Y.-L.; Wang, Y.-H. Susceptible time window and endurable duration of cotton fiber development to high temperature stress. J. Integr. Agric. 2017, 16, 1936–1945. [Google Scholar]
  48. Abdelgadir, H.; Johnson, S.; Van Staden, J. Pollen viability, pollen germination and pollen tube growth in the biofuel seed crop Jatropha curcas (Euphorbiaceae). S. Afr. J. Bot. 2012, 79, 132–139. [Google Scholar] [CrossRef] [Green Version]
  49. Kakani, V.; Reddy, K.; Koti, S.; Wallace, T.; Prasad, P.; Reddy, V.; Zhao, D. Differences in in vitro pollen germination and pollen tube growth of cotton cultivars in response to high temperature. Ann. Bot. 2005, 96, 59–67. [Google Scholar] [CrossRef] [Green Version]
  50. Burke, J.J.; Velten, J.; Oliver, M.J. In vitro analysis of cotton pollen germination. Agron. J. 2004, 96, 359–368. [Google Scholar] [CrossRef]
  51. Hedhly, A.; Hormaza, J.; Herrero, M. The effect of temperature on stigmatic receptivity in sweet cherry (Prunus avium L.). Plant Cell Environ. 2003, 26, 1673–1680. [Google Scholar] [CrossRef] [Green Version]
  52. Hedhly, A.; Hormaza, J.; Herrero, M. The effect of temperature on pollen germination, pollen tube growth, and stigmatic receptivity in peach. Plant Biol. 2005, 7, 476–483. [Google Scholar] [CrossRef] [Green Version]
  53. Snider, J.L.; Oosterhuis, D.M. Heat stress and pollen-pistil interactions. In Flowering and Fruiting in Cotton; Publ. Cotton Foundation: Memphis, TN, USA, 2012; pp. 59–78. [Google Scholar]
  54. Gago, J.; de Menezes Daloso, D.; Figueroa, C.M.; Flexas, J.; Fernie, A.R.; Nikoloski, Z. Relationships of leaf net photosynthesis, stomatal conductance, and mesophyll conductance to primary metabolism: A multispecies meta-analysis approach. Plant Physiol. 2016, 171, 265–279. [Google Scholar] [CrossRef] [PubMed]
  55. Hasanuzzaman, M.; Nahar, K.; Alam, M.; Roychowdhury, R.; Fujita, M. Physiological, biochemical, and molecular mechanisms of heat stress tolerance in plants. Int. J. Mol. Sci. 2013, 14, 9643–9684. [Google Scholar] [CrossRef] [PubMed]
  56. Crafts-Brandner, S.; Law, R. Effect of heat stress on the inhibition and recovery of the ribulose-1, 5-bisphosphate carboxylase/oxygenase activation state. Planta 2000, 212, 67–74. [Google Scholar] [CrossRef] [PubMed]
  57. Schrader, S.; Wise, R.; Wacholtz, W.; Ort, D.R.; Sharkey, T. Thylakoid membrane responses to moderately high leaf temperature in Pima cotton. Plant Cell Environ. 2004, 27, 725–735. [Google Scholar] [CrossRef]
  58. Karademir, E.; Karademir, C.; Sevilmis, U.; Basal, H. Correlations between canopy temperature, chlorophyll content and yield in heat tolerant cotton (Gossypium hirsutum L.) genotypes. Fresenius Environ. Bull. 2018, 27, 5230–5237. [Google Scholar]
  59. Bita, C.; Gerats, T. Plant tolerance to high temperature in a changing environment: Scientific fundamentals and production of heat stress-tolerant crops. Front. Plant Sci. 2013, 4, 273. [Google Scholar] [CrossRef] [Green Version]
  60. Bibi, A.; Oosterhuis, D.; Brown, R.; Gonias, E.; Bourland, F. The physiological response of cotton to high temperatures for germplasm screening. Summ. Ark. Cotton Res. 2003, 521, 87–93. [Google Scholar]
  61. Rahman, H.; Malik, S.A.; Saleem, M. Heat tolerance of upland cotton during the fruiting stage evaluated using cellular membrane thermostability. Field Crops Res. 2004, 85, 149–158. [Google Scholar] [CrossRef]
  62. Khan, N.; Ahmad, I.; Azhar, M.T. Genetic variation in relative cell injury for breeding upland cotton under high temperature stress. Turk. J. Field Crops 2017, 22, 152–159. [Google Scholar] [CrossRef]
  63. Urban, J.; Ingwers, M.; McGuire, M.A.; Teskey, R.O. Stomatal conductance increases with rising temperature. Plant Signal. Behav. 2017, 12, e1356534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Lu, Z.; Percy, R.G.; Qualset, C.O.; Zeiger, E. Stomatal conductance predicts yields in irrigated Pima cotton and bread wheat grown at high temperatures. J. Exp. Bot. 1998, 453–460. [Google Scholar] [CrossRef] [Green Version]
  65. Lu, Z.; Chen, J.; Percy, R.G.; Zeiger, E. Photosynthetic rate, stomatal conductance and leaf area in two cotton species (Gossypium barbadense and Gossypium hirsutum) and their relation with heat resistance and yield. Funct. Plant Biol. 1997, 24, 693–700. [Google Scholar] [CrossRef]
  66. Lu, Z.; Quiñones, M.A.; Zeiger, E. Temperature dependence of guard cell respiration and stomatal conductance co-segregate in an F2 population of Pima cotton. Funct. Plant Biol. 2000, 27, 457–462. [Google Scholar] [CrossRef]
  67. Khan, A.I.; Khan, I.A.; Sadaqat, H.A. Heat tolerance is variable in cotton (Gossypium hirsutum L.) and can be exploited for breeding of better yielding cultivars under high temperature regimes. Pak. J. Bot. 2008, 40, 2053–2058. [Google Scholar]
  68. Rahman, H.; Murtaza, N.; Shah, K.; Qayyum, A.; Ullah, I.; Malik, W. Genetic variation for stomatal conductance in upland cotton as influenced by heat-stressed and non-stressed growing regimes. Acta Agron. Hung. 2008, 56, 11–19. [Google Scholar] [CrossRef]
  69. Suzuki, N.; Mittler, R. Reactive oxygen species and temperature stresses: A delicate balance between signaling and destruction. Physiol. Plant. 2006, 126, 45–51. [Google Scholar] [CrossRef]
  70. Suzuki, N.; Koussevitzky, S.; Mittler, R.; Miller, G. ROS and redox signalling in the response of plants to abiotic stress. Plant Cell Environ. 2012, 35, 259–270. [Google Scholar] [CrossRef]
  71. Considine, M.J.; María Sandalio, L.; Helen Foyer, C. Unravelling how plants benefit from ROS and NO reactions, while resisting oxidative stress. Ann. Bot. 2015, 116, 469–473. [Google Scholar] [CrossRef] [Green Version]
  72. Singh, R.; Singh, S.; Parihar, P.; Mishra, R.K.; Tripathi, D.K.; Singh, V.P.; Chauhan, D.K.; Prasad, S.M. Reactive oxygen species (ROS): Beneficial companions of plants’ developmental processes. Front. Plant Sci. 2016, 7, 1299. [Google Scholar] [CrossRef] [Green Version]
  73. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002, 7, 405–410. [Google Scholar] [CrossRef]
  74. Majeed, S.; Malik, T.A.; Rana, I.A.; Azhar, M.T. Antioxidant and physiological responses of upland cotton accessions grown under high-temperature regimes. Iran. J. Sci. Technol. Trans. A Sci. 2019, 43, 2759–2768. [Google Scholar] [CrossRef]
  75. Sekmen, A.H.; Ozgur, R.; Uzilday, B.; Turkan, I. Reactive oxygen species scavenging capacities of cotton (Gossypium hirsutum) cultivars under combined drought and heat induced oxidative stress. Environ. Exp. Bot. 2014, 99, 141–149. [Google Scholar] [CrossRef]
  76. Gür, A.; Demirel, U.; Özden, M.; Kahraman, A.; Çopur, O. Diurnal gradual heat stress affects antioxidant enzymes, proline accumulation and some physiological components in cotton (Gossypium hirsutum L.). Afr. J. Biotechnol. 2010, 9, 1008–1015. [Google Scholar]
  77. Sarwar, M.; Saleem, M.F.; Ullah, N.; Rizwan, M.; Ali, S.; Shahid, M.R.; Alamri, S.A.; Alyemeni, M.N.; Ahmad, P. Exogenously applied growth regulators protect the cotton crop from heat-induced injury by modulating plant defense mechanism. Sci. Rep. 2018, 8, 17086. [Google Scholar] [CrossRef] [Green Version]
  78. Loka, D.A.; Oosterhuis, D.M. Effect of high night temperatures during anthesis on cotton (Gossypium hirsutum L.) pistil and leaf physiology and biochemistry. Aust. J. Crop Sci. 2016, 10, 741. [Google Scholar] [CrossRef]
  79. Wang, W.; Vinocur, B.; Shoseyov, O.; Altman, A. Role of plant heat-shock proteins and molecular chaperones in the abiotic stress response. Trends Plant Sci. 2004, 9, 244–252. [Google Scholar] [CrossRef]
  80. Kotak, S.; Larkindale, J.; Lee, U.; von Koskull-Döring, P.; Vierling, E.; Scharf, K.-D. Complexity of the heat stress response in plants. Curr. Opin. Plant Biol. 2007, 10, 310–316. [Google Scholar] [CrossRef]
  81. Waters, E.R.; Lee, G.J.; Vierling, E. Evolution, structure and function of the small heat shock proteins in plants. J. Exp. Bot. 1996, 47, 325–338. [Google Scholar] [CrossRef]
  82. Mogk, A.; Schlieker, C.; Friedrich, K.L.; Schönfeld, H.-J.; Vierling, E.; Bukau, B. Refolding of substrates bound to small Hsps relies on a disaggregation reaction mediated most efficiently by ClpB/DnaK. J. Biol. Chem. 2003, 278, 31033–31042. [Google Scholar] [CrossRef] [Green Version]
  83. Van Montfort, R.L.; Basha, E.; Friedrich, K.L.; Slingsby, C.; Vierling, E. Crystal structure and assembly of a eukaryotic small heat shock protein. Nat. Struct. Mol. Biol. 2001, 8, 1025. [Google Scholar] [CrossRef]
  84. Malik, M.K.; Slovin, J.P.; Hwang, C.H.; Zimmerman, J.L. Modified expression of a carrot small heat shock protein gene, Hsp17. 7, results in increased or decreased thermotolerance. Plant J. 1999, 20, 89–99. [Google Scholar] [CrossRef]
  85. Maqbool, A.; Zahur, M.; Irfan, M.; Qaiser, U.; Rashid, B.; Husnain, T. Identification, characterization and expression of drought related alpha-crystalline heat shock protein gene (GHSP26) from Desi cotton. Crop Sci. 2007, 47, 2437–2444. [Google Scholar] [CrossRef]
  86. Lee, G.J.; Vierling, E. A small heat shock protein cooperates with heat shock protein 70 systems to reactivate a heat-denatured protein. Plant Physiol. 2000, 122, 189–198. [Google Scholar] [CrossRef] [Green Version]
  87. Haslbeck, M.; Vierling, E. A first line of stress defense: Small heat shock proteins and their function in protein homeostasis. J. Mol. Biol. 2015, 427, 1537–1548. [Google Scholar] [CrossRef] [Green Version]
  88. Boston, R.S.; Viitanen, P.V.; Vierling, E. Molecular chaperones and protein folding in plants. In Post-Transcriptional Control of Gene Expression in Plants; Springer: Berlin/Heidelberg, Germany, 1996; pp. 191–222. [Google Scholar]
  89. Hartl, F.U.; Bracher, A.; Hayer-Hartl, M. Molecular chaperones in protein folding and proteostasis. Nature 2011, 475, 324. [Google Scholar] [CrossRef]
  90. Sung, D.Y.; Kaplan, F.; Guy, C.L. Plant Hsp70 molecular chaperones: Protein structure, gene family, expression and function. Physiol. Plant. 2001, 113, 443–451. [Google Scholar] [CrossRef]
  91. Scarpeci, T.E.; Zanor, M.I.; Valle, E.M. Investigating the role of plant heat shock proteins during oxidative stress. Plant Signal. Behav. 2008, 3, 856–857. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Li, H.; Liu, S.S.; Yi, C.Y.; Wang, F.; Zhou, J.; Xia, X.J.; Shi, K.; Zhou, Y.H.; Yu, J.Q. Hydrogen peroxide mediates abscisic acid-induced HSP 70 accumulation and heat tolerance in grafted cucumber plants. Plant Cell Environ. 2014, 37, 2768–2780. [Google Scholar] [CrossRef] [PubMed]
  93. Rehman, A.; Atif, R.M.; Qyyum, A.; Du, X.; Hinze, L.; Azhar, M.T. Genome-wide identification and characterization of HSP70 gene family in four species of cotton. Genomics 2020, 112, 4442–4453. [Google Scholar] [CrossRef] [PubMed]
  94. Krishna, P.; Gloor, G. The Hsp90 family of proteins in Arabidopsis thaliana. Cell Stress Chaperones 2001, 6, 238. [Google Scholar] [CrossRef]
  95. Kadota, Y.; Shirasu, K. The HSP90 complex of plants. Biochim. Et Biophys. Acta (BBA)-Mol. Cell Res. 2012, 1823, 689–697. [Google Scholar] [CrossRef] [Green Version]
  96. Sable, A.; Rai, K.M.; Choudhary, A.; Yadav, V.K.; Agarwal, S.K.; Sawant, S.V. Inhibition of heat shock proteins HSP90 and HSP70 induce oxidative stress, suppressing cotton fiber development. Sci. Rep. 2018, 8, 3620. [Google Scholar] [CrossRef] [PubMed]
  97. Zietkiewicz, S.; Krzewska, J.; Liberek, K. Successive and synergistic action of the Hsp70 and Hsp100 chaperones in protein disaggregation. J. Biol. Chem. 2004, 279, 44376–44383. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Winkler, J.; Tyedmers, J.; Bukau, B.; Mogk, A. Hsp70 targets Hsp100 chaperones to substrates for protein disaggregation and prion fragmentation. J. Cell Biol 2012, 198, 387–404. [Google Scholar] [CrossRef] [Green Version]
  99. Hemmingsen, S.M.; Woolford, C.; van der Vies, S.M.; Tilly, K.; Dennis, D.T.; Georgopoulos, C.P.; Hendrix, R.W.; Ellis, R.J. Homologous plant and bacterial proteins chaperone oligomeric protein assembly. Nature 1988, 333, 330. [Google Scholar] [CrossRef]
  100. Apuya, N.R.; Yadegari, R.; Fischer, R.L.; Harada, J.J.; Zimmerman, J.L.; Goldberg, R.B. The Arabidopsis embryo mutant schlepperless has a defect in the chaperonin-60α gene. Plant Physiol. 2001, 126, 717–730. [Google Scholar] [CrossRef] [Green Version]
  101. Ishikawa, A.; Tanaka, H.; Nakai, M.; Asahi, T. Deletion of a chaperonin 60β gene leads to cell death in the Arabidopsis lesion initiation 1 mutant. Plant Cell Physiol. 2003, 44, 255–261. [Google Scholar] [CrossRef]
  102. Zabaleta, E.; Oropeza, A.; Assad, N.; Mandel, A.; Salerno, G.; Herrera-Estrella, L. Antisense expression of chaperonin 60β in transgenic tobacco plants leads to abnormal phenotypes and altered distribution of photoassimilates. Plant J. 1994, 6, 425–432. [Google Scholar] [CrossRef]
  103. Miemyk, J. The 70 kDa stress-related proteins as molecular chaperones. Trends Plant Sci. 1997, 2, 180–187. [Google Scholar] [CrossRef]
  104. Zhang, Y.; Wang, M.; Chen, J.; Rong, J.; Ding, M. Genome-wide analysis of HSP70 superfamily in Gossypium raimondii and the expression of orthologs in Gossypium hirsutum. Yi Chuan Hered. 2014, 36, 921–933. [Google Scholar]
  105. Young, J.C.; Moarefi, I.; Hartl, F.U. Hsp90: A specialized but essential protein-folding tool. J. Cell Biol. 2001, 154, 267. [Google Scholar] [CrossRef] [PubMed]
  106. Buchner, J. Hsp90 & Co.—A holding for folding. Trends Biochem. Sci. 1999, 24, 136–141. [Google Scholar]
  107. Zhang, Z.; Quick, M.K.; Kanelakis, K.C.; Gijzen, M.; Krishna, P. Characterization of a plant homolog of hop, a cochaperone of hsp90. Plant Physiol. 2003, 131, 525–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Latterich, M.; Patel, S. The AAA team: Related ATPases with diverse functions. Trends Cell Biol. 1998, 8, 65–71. [Google Scholar] [CrossRef]
  109. Lee, U.; Rioflorido, I.; Hong, S.W.; Larkindale, J.; Waters, E.R.; Vierling, E. The Arabidopsis ClpB/Hsp100 family of proteins: Chaperones for stress and chloroplast development. Plant J. 2007, 49, 115–127. [Google Scholar] [CrossRef] [Green Version]
  110. Mishra, R.C.; Grover, A. ClpB/Hsp100 proteins and heat stress tolerance in plants. Crit. Rev. Biotechnol. 2016, 36, 862–874. [Google Scholar] [CrossRef]
  111. Barrera-Figueroa, B.E.; Gao, L.; Wu, Z.; Zhou, X.; Zhu, J.; Jin, H.; Liu, R.; Zhu, J.-K. High throughput sequencing reveals novel and abiotic stress-regulated microRNAs in the inflorescences of rice. BMC Plant Biol. 2012, 12, 132. [Google Scholar] [CrossRef] [Green Version]
  112. Barakat, A.; Sriram, A.; Park, J.; Zhebentyayeva, T.; Main, D.; Abbott, A. Genome wide identification of chilling responsive microRNAs in Prunus persica. BMC Genom. 2012, 13, 481. [Google Scholar] [CrossRef] [Green Version]
  113. Huang, J.; Yang, M.; Zhang, X. The function of small RNAs in plant biotic stress response. J. Integr. Plant Biol. 2016, 58, 312–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. Wang, J.-W.; Czech, B.; Weigel, D. miR156-regulated SPL transcription factors define an endogenous flowering pathway in Arabidopsis thaliana. Cell 2009, 138, 738–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Chen, G.; Li, J.; Liu, Y.; Zhang, Q.; Gao, Y.; Fang, K.; Cao, Q.; Qin, L.; Xing, Y. Roles of the GA-mediated SPL Gene Family and miR156 in the Floral Development of Chinese Chestnut (Castanea mollissima). Int. J. Mol. Sci. 2019, 20, 1577. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Liu, N.; Tu, L.; Wang, L.; Hu, H.; Xu, J.; Zhang, X. MicroRNA 157-targeted SPL genes regulate floral organ size and ovule production in cotton. BMC Plant Biol. 2017, 17, 7. [Google Scholar] [CrossRef] [Green Version]
  117. Yu, X.; Wang, H.; Lu, Y.; de Ruiter, M.; Cariaso, M.; Prins, M.; van Tunen, A.; He, Y. Identification of conserved and novel microRNAs that are responsive to heat stress in Brassica rapa. J. Exp. Bot. 2011, 63, 1025–1038. [Google Scholar] [CrossRef]
  118. Song, C.; Fang, J.; Li, X.; Liu, H.; Chao, C.T. Identification and characterization of 27 conserved microRNAs in citrus. Planta 2009, 230, 671–685. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Stief, A.; Altmann, S.; Hoffmann, K.; Pant, B.D.; Scheible, W.-R.; Bäurle, I. Arabidopsis miR156 regulates tolerance to recurring environmental stress through SPL transcription factors. Plant Cell 2014, 26, 1792–1807. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Wang, Y.; Sun, F.; Cao, H.; Peng, H.; Ni, Z.; Sun, Q.; Yao, Y. TamiR159 directed wheat TaGAMYB cleavage and its involvement in anther development and heat response. PLoS ONE 2012, 7, e48445. [Google Scholar] [CrossRef] [Green Version]
  121. Ding, Y.; Ma, Y.; Liu, N.; Xu, J.; Hu, Q.; Li, Y.; Wu, Y.; Xie, S.; Zhu, L.; Min, L. micro RNA s involved in auxin signalling modulate male sterility under high-temperature stress in cotton (Gossypium hirsutum). Plant J. 2017, 91, 977–994. [Google Scholar] [CrossRef] [Green Version]
  122. Wang, M.; Wang, Q.; Zhang, B. Response of miRNAs and their targets to salt and drought stresses in cotton (Gossypium hirsutum L.). Gene 2013, 530, 26–32. [Google Scholar] [CrossRef]
  123. Sailaja, B.; Anjum, N.; Prasanth, V.V.; Sarla, N.; Subrahmanyam, D.; Voleti, S.; Viraktamath, B.; Mangrauthia, S.K. Comparative study of susceptible and tolerant genotype reveals efficient recovery and root system contributes to heat stress tolerance in rice. Plant Mol. Biol. Rep. 2014, 32, 1228–1240. [Google Scholar] [CrossRef]
  124. Ruan, M.-B.; Zhao, Y.-T.; Meng, Z.-H.; Wang, X.-J.; Yang, W.-C. Conserved miRNA analysis in Gossypium hirsutum through small RNA sequencing. Genomics 2009, 94, 263–268. [Google Scholar] [CrossRef] [Green Version]
  125. Boopathi, M.; Sathish, S.; Kavitha, P.; Dachinamoorthy, P.; Ravikesavan, R. Comparative miRNAome analysis revealed numerous conserved and novel drought responsive miRNAs in cotton (Gossypium spp.). Cotton Genom. Genet. 2016. [Google Scholar] [CrossRef] [Green Version]
  126. Fang, Y.; Xie, K.; Xiong, L. Conserved miR164-targeted NAC genes negatively regulate drought resistance in rice. J. Exp. Bot. 2014, 65, 2119–2135. [Google Scholar] [CrossRef] [Green Version]
  127. Kumar, R.R.; Pathak, H.; Sharma, S.K.; Kala, Y.K.; Nirjal, M.K.; Singh, G.P.; Goswami, S.; Rai, R.D. Novel and conserved heat-responsive microRNAs in wheat (Triticum aestivum L.). Funct. Integr. Genom. 2015, 15, 323–348. [Google Scholar] [CrossRef] [PubMed]
  128. Matthews, C.; Arshad, M.; Hannoufa, A. Alfalfa response to heat stress is modulated by microRNA156. Physiol. Plant. 2019, 165, 830–842. [Google Scholar] [CrossRef] [PubMed]
  129. Zhao, J.; He, Q.; Chen, G.; Wang, L.; Jin, B. Regulation of non-coding RNAs in heat stress responses of plants. Front. Plant Sci. 2016, 7, 1213. [Google Scholar] [CrossRef] [Green Version]
  130. Zhang, B.; Liu, J.; Yang, Z.E.; Chen, E.Y.; Zhang, C.J.; Zhang, X.Y.; Li, F.G. Genome-wide analysis of GRAS transcription factor gene family in Gossypium hirsutum L. BMC Genom. 2018, 19, 348. [Google Scholar] [CrossRef]
  131. Aukerman, M.J.; Sakai, H. Regulation of flowering time and floral organ identity by a microRNA and its APETALA2-like target genes. Plant Cell 2003, 15, 2730–2741. [Google Scholar] [CrossRef] [Green Version]
  132. François, L.; Verdenaud, M.; Fu, X.; Ruleman, D.; Dubois, A.; Vandenbussche, M.; Bendahmane, A.; Raymond, O.; Just, J.; Bendahmane, M. A miR172 target-deficient AP2-like gene correlates with the double flower phenotype in roses. Sci. Rep. 2018, 8, 12912. [Google Scholar] [CrossRef] [Green Version]
  133. Wu, G.; Park, M.Y.; Conway, S.R.; Wang, J.-W.; Weigel, D.; Poethig, R.S. The sequential action of miR156 and miR172 regulates developmental timing in Arabidopsis. Cell 2009, 138, 750–759. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Zhu, Q.-H.; Upadhyaya, N.M.; Gubler, F.; Helliwell, C.A. Over-expression of miR172 causes loss of spikelet determinacy and floral organ abnormalities in rice (Oryza sativa). BMC Plant Biol. 2009, 9, 149. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Yin, Z.; Han, X.; Li, Y.; Wang, J.; Wang, D.; Wang, S.; Fu, X.; Ye, W. Comparative analysis of cotton small RNAs and their target genes in response to salt stress. Genes 2017, 8, 369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Wójcik, A.M.; Gaj, M.D. miR393 contributes to the embryogenic transition induced in vitro in Arabidopsis via the modification of the tissue sensitivity to auxin treatment. Planta 2016, 244, 231–243. [Google Scholar] [CrossRef] [Green Version]
  137. Jeong, D.-H.; Green, P.J. The role of rice microRNAs in abiotic stress responses. J. Plant Biol. 2013, 56, 187–197. [Google Scholar] [CrossRef]
  138. Xie, F.; Wang, Q.; Sun, R.; Zhang, B. Deep sequencing reveals important roles of microRNAs in response to drought and salinity stress in cotton. J. Exp. Bot. 2014, 66, 789–804. [Google Scholar] [CrossRef] [Green Version]
  139. Wang, H.; Huang, J.; Lai, Z.; Xue, Y. F-box proteins in flowering plants. Chin. Sci. Bull. 2002, 47, 1497–1501. [Google Scholar] [CrossRef] [Green Version]
  140. Xu, G.; Ma, H.; Nei, M.; Kong, H. Evolution of F-box genes in plants: Different modes of sequence divergence and their relationships with functional diversification. Proc. Natl. Acad. Sci. USA 2009, 106, 835–840. [Google Scholar] [CrossRef] [Green Version]
  141. Song, J.B.; Gao, S.; Sun, D.; Li, H.; Shu, X.X.; Yang, Z.M. miR394 and LCR are involved in Arabidopsis salt and drought stress responses in an abscisic acid-dependent manner. BMC Plant Biol. 2013, 13, 210. [Google Scholar] [CrossRef] [Green Version]
  142. Liu, Q.; Yang, T.; Yu, T.; Zhang, S.; Mao, X.; Zhao, J.; Wang, X.; Dong, J.; Liu, B. Integrating small RNA sequencing with QTL mapping for identification of miRNAs and their target genes associated with heat tolerance at the flowering stage in rice. Front. Plant Sci. 2017, 8, 43. [Google Scholar] [CrossRef]
  143. Zhang, B. MicroRNA: A new target for improving plant tolerance to abiotic stress. J. Exp. Bot. 2015, 66, 1749–1761. [Google Scholar] [CrossRef]
  144. Majeed, S.; Chaudhary, M.T.; Hulse-Kemp, A.M.; Azhar, M.T. Introduction: Crop Wild Relatives in Plant Breeding. In Wild Germplasm for Genetic Improvement in Crop Plants; Elsevier: Amsterdam, The Netherlands, 2021; pp. 1–18. [Google Scholar]
  145. Raza, A.; Tabassum, J.; Kudapa, H.; Varshney, R.K. Can omics deliver temperature resilient ready-to-grow crops? Crit. Rev. Biotechnol. 2021, 1–24. [Google Scholar] [CrossRef]
  146. Percy, R.G. Comparison of bulk F2 performance testing and pedigree selection in thirty Pima cotton populations. J. Cotton Sci. 2003, 7, 123. [Google Scholar]
  147. Tokatlidis, I.; Tsikrikoni, C.; Lithourgidis, A.; Tsialtas, J.; Tzantarmas, C. Intra-cultivar variation in cotton: Response to single-plant yield selection at low density. J. Agric. Sci. 2011, 149, 197–204. [Google Scholar] [CrossRef]
  148. Mubarik, M.S.; Ma, C.; Majeed, S.; Du, X.; Azhar, M.T. Revamping of Cotton Breeding Programs for Efficient Use of Genetic Resources under Changing Climate. Agronomy 2020, 10, 1190. [Google Scholar] [CrossRef]
  149. Badigannavar, A.; Myers, G.O.; Jones, D.C. Molecular diversity revealed by AFLP markers in upland cotton genotypes. J. Crop Improv. 2012, 26, 627–640. [Google Scholar] [CrossRef]
  150. Mohamed, H.; Abdel-Hamid, A. Molecular and biochemical studies for heat tolerance on four cotton genotypes. Rom. Biotechnol. Lett. 2013, 18, 8823–8831. [Google Scholar]
  151. Du, L.; Cai, C.; Wu, S.; Zhang, F.; Hou, S.; Guo, W. Evaluation and exploration of favorable QTL alleles for salt stress related traits in cotton cultivars (G. hirsutum L.). PLoS ONE 2016, 11, e0151076. [Google Scholar] [CrossRef] [Green Version]
  152. Majeed, S.; Rana, I.A.; Atif, R.M.; Zulfiqar, A.; Hinze, L.; Azhar, M.T. Role of SNPs in determining QTLs for major traits in cotton. J. Cotton Res. 2019, 2, 5. [Google Scholar] [CrossRef] [Green Version]
  153. Demirel, U.; Gür, A.; Can, N.; Memon, A. Identification of heat responsive genes in cotton. Biol. Plant. 2014, 58, 515–523. [Google Scholar] [CrossRef]
  154. Zhang, J.; Srivastava, V.; Stewart, J.M.; Underwood, J. Heat-tolerance in Cotton Is Correlated with Induced Overexpression of Heat-Shock Factors, Heat-Shock Proteins, and General Stress Response Genes. J. Cotton Sci. 2016, 20, 253–262. [Google Scholar]
  155. Shaheen, T.; Asif, M.; Zafar, Y. Single nucleotide polymorphism analysis of MT-SHSP gene of Gossypium arboreum and its relationship with other diploid cotton genomes, G. hirsutum and Arabidopsis thaliana. Pak. J. Bot. 2009, 41, 117–183. [Google Scholar]
  156. Ma, W.; Zhao, T.; Li, J.; Liu, B.; Fang, L.; Hu, Y.; Zhang, T. Identification and characterization of the GhHsp20 gene family in Gossypium hirsutum. Sci. Rep. 2016, 6, 32517. [Google Scholar] [CrossRef] [Green Version]
  157. Islam, M.S.; Fang, D.D.; Jenkins, J.N.; Guo, J.; McCarty, J.C.; Jones, D.C. Evaluation of genomic selection methods for predicting fiber quality traits in Upland cotton. Mol. Genet. Genom. 2020, 295, 67–79. [Google Scholar] [CrossRef] [PubMed]
  158. Baytar, A.A.; Peynircioğlu, C.; Sezener, V.; Basal, H.; Frary, A.; Frary, A.; Doğanlar, S. Genome-wide association mapping of yield components and drought tolerance-related traits in cotton. Mol. Breed. 2018, 38, 74. [Google Scholar] [CrossRef]
  159. Su, J.; Pang, C.; Wei, H.; Li, L.; Liang, B.; Wang, C.; Song, M.; Wang, H.; Zhao, S.; Jia, X. Identification of favorable SNP alleles and candidate genes for traits related to early maturity via GWAS in upland cotton. BMC Genom. 2016, 17, 687. [Google Scholar] [CrossRef] [Green Version]
  160. Batcho, A.A.; Sarwar, M.B.; Rashid, B.; Hassan, S.; Husnain, T. Heat shock protein gene identified from Agave sisalana (As HSP70) confers heat stress tolerance in transgenic cotton (Gossypium hirsutum). Theor. Exp. Plant Physiol. 2021, 33, 141–156. [Google Scholar] [CrossRef]
  161. Esmaeili, N.; Cai, Y.; Tang, F.; Zhu, X.; Smith, J.; Mishra, N.; Hequet, E.; Ritchie, G.; Jones, D.; Shen, G. Towards doubling fibre yield for cotton in the semiarid agricultural area by increasing tolerance to drought, heat and salinity simultaneously. Plant Biotechnol. J. 2021, 19, 462. [Google Scholar] [CrossRef]
  162. Burke, J.J.; Chen, J. Enhancement of reproductive heat tolerance in plants. PLoS ONE 2015, 10, e0122933. [Google Scholar] [CrossRef] [Green Version]
  163. Mishra, N.; Sun, L.; Zhu, X.; Smith, J.; Prakash Srivastava, A.; Yang, X.; Pehlivan, N.; Esmaeili, N.; Luo, H.; Shen, G. Overexpression of the rice SUMO E3 ligase gene OsSIZ1 in cotton enhances drought and heat tolerance, and substantially improves fiber yields in the field under reduced irrigation and rainfed conditions. Plant Cell Physiol. 2017, 58, 735–746. [Google Scholar] [CrossRef]
  164. Hozain, M.d.; Abdelmageed, H.; Lee, J.; Kang, M.; Fokar, M.; Allen, R.D.; Holaday, A.S. Expression of AtSAP5 in cotton up-regulates putative stress-responsive genes and improves the tolerance to rapidly developing water deficit and moderate heat stress. J. Plant Physiol. 2012, 169, 1261–1270. [Google Scholar] [CrossRef]
  165. Ali, S.; Rizwan, M.; Arif, M.S.; Ahmad, R.; Hasanuzzaman, M.; Ali, B.; Hussain, A. Approaches in enhancing thermotolerance in plants: An updated review. J. Plant Growth Regul. 2020, 39, 456–480. [Google Scholar] [CrossRef]
  166. Meriç, S.; Ayan, A.; Atak, Ç. Molecular Abiotic Stress Tolerans Strategies: From Genetic Engineering to Genome Editing Era. In Abiotic Stress in Plants; IntechOpen: London, UK, 2020. [Google Scholar]
  167. Mubarik, M.S.; Khan, S.H.; Ahmad, A.; Khan, Z.; Sajjad, M.; Khan, I.A. Disruption of Phytoene Desaturase Gene using Transient Expression of Cas9: gRNA Complex. Int. J. Agric. Biol. 2016, 18. [Google Scholar] [CrossRef]
  168. Bhatnagar-Mathur, P.; Vadez, V.; Sharma, K.K. Transgenic approaches for abiotic stress tolerance in plants: Retrospect and prospects. Plant Cell Rep. 2008, 27, 411–424. [Google Scholar] [CrossRef]
  169. Ruiz-Vera, U.M.; Siebers, M.H.; Drag, D.W.; Ort, D.R.; Bernacchi, C.J. Canopy warming caused photosynthetic acclimation and reduced seed yield in maize grown at ambient and elevated [CO2]. Glob. Chang. Biol. 2015, 21, 4237–4249. [Google Scholar] [CrossRef] [PubMed]
  170. Qiu, Z.; Kang, S.; He, L.; Zhao, J.; Zhang, S.; Hu, J.; Zeng, D.; Zhang, G.; Dong, G.; Gao, Z. The newly identified heat-stress sensitive albino 1 gene affects chloroplast development in rice. Plant Sci. 2018, 267, 168–179. [Google Scholar] [CrossRef] [PubMed]
  171. Klap, C.; Yeshayahou, E.; Bolger, A.M.; Arazi, T.; Gupta, S.K.; Shabtai, S.; Usadel, B.; Salts, Y.; Barg, R. Tomato facultative parthenocarpy results from Sl AGAMOUS-LIKE 6 loss of function. Plant Biotechnol. J. 2017, 15, 634–647. [Google Scholar] [CrossRef]
  172. Li, J.; Zhang, H.; Si, X.; Tian, Y.; Chen, K.; Liu, J.; Chen, H.; Gao, C. Generation of thermosensitive male-sterile maize by targeted knockout of the ZmTMS5 gene. J. Genet. Genom. Yi Chuan Xue Bao 2017, 44, 465–468. [Google Scholar] [CrossRef]
  173. Mubarik, M.S.; Khan, S.H.; Sajjad, M.; Raza, A.; Hafeez, M.B.; Yasmeen, T.; Rizwan, M.; Ali, S.; Arif, M.S. A manipulative interplay between positive and negative regulators of phytohormones: A way forward for improving drought tolerance in plants. Physiol. Plant. 2021. [Google Scholar] [CrossRef] [PubMed]
  174. Yin, W.; Xiao, Y.; Niu, M.; Meng, W.; Li, L.; Zhang, X.; Liu, D.; Zhang, G.; Qian, Y.; Sun, Z. ARGONAUTE2 enhances grain length and salt tolerance by activating BIG GRAIN3 to modulate cytokinin distribution in rice. Plant Cell 2020, 32, 2292–2306. [Google Scholar] [CrossRef]
  175. Zhai, N.; Jia, H.; Liu, D.; Liu, S.; Ma, M.; Guo, X.; Li, H. GhMAP3K65, a cotton Raf-like MAP3K gene, enhances susceptibility to pathogen infection and heat stress by negatively modulating growth and development in transgenic Nicotiana benthamiana. Int. J. Mol. Sci. 2017, 18, 2462. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Salient adverse effects of high temperature stress on cotton plants.
Figure 1. Salient adverse effects of high temperature stress on cotton plants.
Agronomy 11 01825 g001
Figure 2. Effects of high temperature stress on reproductive phases of plant.
Figure 2. Effects of high temperature stress on reproductive phases of plant.
Agronomy 11 01825 g002
Figure 3. Schematic diagram of ROS production in plants.
Figure 3. Schematic diagram of ROS production in plants.
Agronomy 11 01825 g003
Table 1. Characterization of various groups of HSP in plants.
Table 1. Characterization of various groups of HSP in plants.
Group
(Sub-Families)
Representative MembersIntracellular LocalizationMajor RoleReference
sHSP Stabilization of non-native proteins and prevent aggregation[86,87]
I Hsp17.6 Cytosol
IIHsp17.9Cytosol
III Hsp21, Hsp26.2 5Chloroplast
IVHsp22Endoplasmic reticulum
VHsp23 5Mitochondria
VIHsp22.3 Membrane
HSP60 Folding of proteins[88,89]
Group ICpn60 2Chloroplast, mitochondria
Group II CCT3Cytosol
HSP70 Protein import, signal transduction, transcriptional activation, assist refolding and prevent aggregation of proteins [90,91,92,93]
DnaK Hsp/Hsc70Cytosol
Hsp70Chloroplast, mitochondria
Bip 1Endoplasmic reticulum
Hsp110/SSE Hsp91Cytosol
HSP90AtHsp90-1CytosolFacilitate in genetic buffering and maturation of signaling molecules[94,95,96]
AtHsp90-5Chloroplast
AtHsp90-6Mitochondria
AtHsp90-7Endoplasmic reticulum
HSP100 Unfolding and disaggregation of proteins[97,98]
Class IClpB, ClpA/C, ClpDCytosol, mitochondria
Class II ClpM, ClpN, ClpX, ClpYChloroplast
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Majeed, S.; Rana, I.A.; Mubarik, M.S.; Atif, R.M.; Yang, S.-H.; Chung, G.; Jia, Y.; Du, X.; Hinze, L.; Azhar, M.T. Heat Stress in Cotton: A Review on Predicted and Unpredicted Growth-Yield Anomalies and Mitigating Breeding Strategies. Agronomy 2021, 11, 1825. https://doi.org/10.3390/agronomy11091825

AMA Style

Majeed S, Rana IA, Mubarik MS, Atif RM, Yang S-H, Chung G, Jia Y, Du X, Hinze L, Azhar MT. Heat Stress in Cotton: A Review on Predicted and Unpredicted Growth-Yield Anomalies and Mitigating Breeding Strategies. Agronomy. 2021; 11(9):1825. https://doi.org/10.3390/agronomy11091825

Chicago/Turabian Style

Majeed, Sajid, Iqrar Ahmad Rana, Muhammad Salman Mubarik, Rana Muhammad Atif, Seung-Hwan Yang, Gyuhwa Chung, Yinhua Jia, Xiongming Du, Lori Hinze, and Muhammad Tehseen Azhar. 2021. "Heat Stress in Cotton: A Review on Predicted and Unpredicted Growth-Yield Anomalies and Mitigating Breeding Strategies" Agronomy 11, no. 9: 1825. https://doi.org/10.3390/agronomy11091825

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