Skip to main content

Enhancing cold and drought tolerance in cotton: a protective role of SikCOR413PM1

Abstract

The present study explored the potential role of cold-regulated plasma membrane protein COR413PM1 isolated from Saussurea involucrata (Matsum. & Koidz)(SikCOR413PM1), in enhancing cotton (Gossypium hirsutum) tolerance to cold and drought stresses through transgenic methods. Under cold and drought stresses, the survival rate and the fresh and dry weights of the SikCOR413PM1-overexpressing lines were higher than those of the wild-type plants, and the degree of leaf withering was much lower. Besides, overexpressing SikCOR413PM1 overexpression increased the relative water content, reduced malondialdehyde content and relative conductivity, and elevated proline and soluble sugar levels in cotton seedlings. These findings suggest that SikCOR413PM1 minimizes cell membrane damage and boosts plant stability under challenging conditions. Additionally, overexpression of this gene upregulated antioxidant enzyme-related genes in cotton seedlings, resulting in enhanced antioxidant enzyme activity, lowered peroxide content, and reduced oxidative stress. SikCOR413PM1 overexpression also modulated the expression of stress-related genes (GhDREB1A, GhDREB1B, GhDREB1C, GhERF2, GhNAC3, and GhRD22). In field trials, the transgenic cotton plants overexpressing SikCOR413PM1 displayed high yields and increased environmental tolerance. Our study thus demonstrates the role of SikCOR413PM1 in regulating stress-related genes, osmotic adjustment factors, and peroxide content while preserving cell membrane stability and improving cold and drought tolerance in cotton.

Peer Review reports

Background

Genomic breeding strategies can be used to design future crops. Improved varieties generated via such technologies will produce high yields with minimal agronomic inputs and will be better adapted to climate change [1]. During growth and development, plants are often challenged by abiotic stress factors, such as cold, drought, and salinity [2]. Among these, cold and drought are the most common and greatly impact cotton production and yields [3, 4]. Cold stress causes plant dwarfing, leaf decolorization, and reduced flowering [5]. It also affects pollen viability and pollen tube growth [6], reducing pollination and seed setting [7, 8]. On the other hand, drought stress causes stomatal closure, limits water transpiration [9], increases heat susceptibility, and minimizes nutrient absorption and photosynthesis [10] in plant, thereby slowing growth and reducing yield [11,12,13]. However, exposure to a small degree of abiotic stress in advance can improve the plant’s adaptability to subsequent stress [14].

Plants have evolved various regulatory mechanisms to cope with the changes in environmental conditions. Under normal conditions, plants produce a relatively low reactive oxygen species (ROS) and balance ROS scavenging and production [15]. However, stress can lead to an increase in ROS production [16]. Excessive ROS accumulation causes cell oxidative damage, inhibiting plant physiological processes [17, 18]. Superoxide (O2) and hydrogen peroxide (H2O2) are the major ROS that act as signal transduction molecules mediating multiple stress tolerance [19]. Among these, H2O2 triggers the plant antioxidant system through abscisic acid (ABA), Melatonin (MET) and brassinosteroids (BRs), regulates cold stress, and improves cold tolerance [20,21,22,23,24,25]. Numerous studies have shown that the antioxidant defense mechanism of ROS scavenging is essential in resisting various stresses [26,27,28].

Plant responses to cold stress involve multiple levels of regulation, including metabolic changes, such as the accumulation of sugars and amino acids, and the CBF/DREB pathway [29, 30]. CBF proteins bind to the c-repeat (CRT) cis-elements and activate cold-responsive (COR) gene transcription to improve cold tolerance [21, 31,32,33,34]. Thus, the CBF transcription factor is a positive regulator of cold signals. Numerous studies have demonstrated that the expression of the COR gene in plants is positively correlated with cold resistance [35,36,37,38].

The COR413 family encodes plant-specific multi-transmembrane proteins [39]. Based on the predicted intracellular localization, the COR413 protein family is divided into two subclasses: chloroplast COR413-TM/IM (thylakoid membrane/intima) and plasma membrane COR413-PM [39]. Studies have demonstrated the role of COR413 in improving plant stress tolerance. For example, AtCOR413IM of Arabidopsis thaliana stabilizes the chloroplast membrane under cold stress [40]. Besides, the overexpression of COR413-TM1 in rice enhanced drought tolerance [41], and that of tomato chloroplast COR413 (SlCOR413IM1) in tobacco improved drought tolerance [42]. Similarly, COR413-PM, located in the plasma membrane, is also known to regulate plant stress tolerance. Overexpression of SikCOR413PM1 isolated from Saussurea involucrata (Matsum. & Koidz) in tobacco enhanced the tolerance of transgenic tobacco to cold and drought stress [43]. The overexpression of PsCOR413PM2 in Arabidopsis increased osmotic stress and cold tolerance [44]. LeCOR413PM2 overexpression in the tomato also improved cold tolerance [45].

Cotton is one of the most important agricultural products in the world and is mainly grown in arid regions, such as tropical and subtropical regions [46]. Cotton is grown in over 80 countries [47] and is an important cash crop in China. Xinjiang, the main production area of cotton in China, is rich in land resources; however, the distribution of water resources in this region is extremely uneven, and the temperature is relatively low in early spring during the growth of cotton seedlings [48]. Therefore, much cotton is subjected to drought and cold stress at the seedling stage. Cotton fiber yield and quality are often not guaranteed after experiencing abiotic stresses such as cold and drought [49, 50]. Besides, the simultaneous occurrence of two or more abiotic stresses causes a much greater yield loss than a single stress [51]. Therefore, there is an urgent need to develop cotton cultivars with multiple stress tolerance and meet the increasing demand of people. Studies have demonstrated the use of transgenic methods to improve cotton tolerance. For example, Basso et al. (2021) overexpressed the CaHB12 transcription factor in cotton and increased its drought tolerance [52]. Overexpression of GthCBF4 by Liu et al. (2021) significantly improved the cold stress tolerance of cotton [53]. Zhang et al. (2021) found that overexpressing GhEXLB2 in cotton enhanced drought tolerance at the germination, seedling, and flowering stages [54]. Similarly, overexpressing AmCBF1 [55] and GhGLK1 [56] increased cold and drought tolerance. On the other hand, GhORP_A02 gene silencing enhanced the drought tolerance of cotton [57]. However, there is still a need to develop tolerant cotton genotypes to overcome cold and drought stress and maintain their yield and fiber quality.

Previously, overexpression of the SikCOR413PM1 gene isolated from S. involucrata (Matsum. & Koidz) [58] improved stress resistance in tobacco [43]. However, most conventional methods to develop the stress tolerance require laborious characterization of the progenies of multiple generations derived from time-consuming genetic crosses [59]. In such cases, phenotypic analysis helps investigate gene function. Therefore, the present study explored the role of SikCOR413PM1 overexpression in cotton based on the Liu et al. (2017) approach [60]. The study showed stronger stress tolerance of SikCOR413PM1-overexpressing cotton plants than wild-type (WT) plants in greenhouse and field trials. These findings provide a foundation for further research on SikCOR413PM1 and its application in developing new cotton varieties.

Results

Identification of transgenic cotton plants overexpressing SikCOR431PM1

To investigate the influence of SikCOR431PM1 on cold and drought tolerance, we integrated this gene into cotton following the pollen tube channel method and generated transgenic plants. Ten transgenic cotton lines named OE-1 to 10 (We obtained ten transgenic cotton plants, which were defined as overexpression 1–10, or OE1-10 for short) were screened and analyzed by qRT-PCR (Fig. 1A). Among these lines, OE-4, OE-5, and OE-7 exhibited higher relative expression levels of SikCOR431PM1 than the WT plants. Further analysis based on RT-PCR revealed expression levels consistent with qRT-PCR (Fig. 1B). Therefore, OE-4, OE-5, and OE-7 were selected for studying the gene function.The expression level at 0h (control time point) was defined as 1.0.

Fig. 1
figure 1

Expression levels of SikCOR413PM1 in WT and transgenic cotton. (A) Quantitative real-time PCR and (B) semi-quantitative PCR show SikCOR413PM1 expression in the WT and transgenic (OE-4, OE-5, and OE-7) cotton lines

Growth of SikCOR413PM1- overexpressing cotton seedlings under cold and drought stresses

The WT and transgenic cotton seedlings were exposed to cold (4 °C) and drought (20% PEG6000) stresses to investigate the role of SikCOR413PM1. The WT and transgenic lines grew well under normal conditions (25 °C). However, after exposure to 4 °C for 24 h, the leaves of WT cotton seedlings showed severe wilting, and most became necrotic and damaged. In contrast, the leaves of transgenic lines showed certain vitality and less wilting. After 48 h the leaves of wild-type plants withered much more than transgenic strain. After 72 h the whole wilted leaves of the wild-type plant turned brown, while the transgenic strain only showed yellowing and partial wilting of the leaves. (Fig. 2A). After treatment with 20% PEG6000 for 10days, the leaves of WT plants exhibited noticeable wilting and curling, while those of the transgenic lines showed less wilting. After 15 days, the WT lines exhibited more wilting and withering than the transgenic lines. After 20 days, a few WT plants exhibited even more severe growth retardation and leaf wilting [61, 62], whereas the transgenic lines showed less leaf wilting (Fig. 2B).

To further clarify whether SikCOR413PM1 overexpression can improve the cytoprotection of cotton seedlings under cold and drought stress, we measured the survival rates of these plants (OE-4, OE-5, and OE-7) under cold and 20% PEG6000 stress. Under normal conditions, both WT and transgenic plants exhibited 100% survival. However, after exposure to 4 °C for 72 h, the survival rates of both WT and the transgenic lines decreased; the survival rate of WT decreased to 23.3%, while that of the transgenic lines OE-4, OE-5, and OE-7 decreased to 65.6%, 73.3%, and 80%, respectively. Similarly, after treatment with 20% PEG6000 for 20d, the survival rates of the WT and the OE-4, OE-5, and OE-7 transgenic lines were reduced to 33.3%, 61.1%, 53.3%, and 68.9%, respectively. These observations indicate a higher survival rate of SikCOR413PM1-overexpressing plants than the WT under cold and drought stresses (Fig. 2C).

Under normal conditions, no significant difference was observed in the fresh and dry weights between the WT and transgenic lines. However, after exposure to cold stress for 72 h, the fresh weight of OE-4, OE-5, and OE-7 was 23.6%, 36.6%, and 41.3% higher than that of the WT, respectively (Fig. 2D), while the dry weight was 23.7%, 24.1%, and 28.1% higher (Fig. 2E). Similarly, exposure to 20% PEG6000 stress for 20d resulted in a decline in the fresh and dry weights of both the WT and transgenic plants (OE-4, OE-5, and OE-7) compared with the untreated control. Nevertheless, OE-4, OE-5, and OE-7 lines exhibited 56.0%, 50.9%, and 72.5% higher fresh weight than the WT, and 56.5%, 44.7%, and 63.5% higher dry weight. These results indicate that SikCOR413PM1 overexpression enhances the tolerance of cotton seedlings to cold and drought stress.

Fig. 2
figure 2

Phenotypic analysis of 20-day-old WT and SikCOR413PM1-overexpressing cotton seedlings under cold and drought stress. (A) Phenotype of WT and transgenic (OE-4, OE-5, and OE-7) lines under 4 °C for 24, 48, and 72 h; (B) Phenotype of WT and transgenic (OE-4, OE-5, and OE-7) lines under 20% PGE6000 stress for 10, 15, and 20 d; (C) Survival rate; (D) fresh weight and (E) dry weight of the plants under stress. The data shown are the average of three independent biological replicates. Bars represent standard deviations (SDs). * and ** indicate significant differences relative to the WT plants at P < 0.05 and P < 0.01, respectively. (In the cold and drought treatment figure (Fig. 2A&B), the upper left corner of the pot is the wild type, and the upper right, lower left and lower right corners are transgenic (OE-4, OE-5, and OE-7) lines, respectively)

Overexpression of SikCOR413PM1 enhanced cold and drought tolerance of cotton seedlings

The WT and transgenic cotton plants exhibited robust growth after 70 days at room temperature (25 °C) without treatment. However, after 24 h of exposure to 4 °C, the top leaves of the WT plants showed slight wilting, while those of the transgenics showed no significant change. After 48 h of exposure, the WT plants exhibited severe wilting of leaves at the top, while the transgenic plants showed only slight wilting and drooping of leaves (Fig. 3A). Similarly, after 14 days of exposure to drought, the WT plants showed apparent signs of wilting, with loose and yellow leaves near the base. In contrast, the leaves of transgenic cotton showed less wilting. After 21 days of exposure to drought, the WT plants exhibited severe growth retardation and leaf wilting and shedding. On the other hand, the trangenic plants’ leaves did not fall off; moreover, the transgenic stems remained full throughout the treatment period (Fig. 3B).

Under normal conditions, the WT and the transgenic plants showed no significant difference in average fresh and dry weights. However, after exposure to cold, the WT plants had 19.4%, 24.2%, and 25.5% lower fresh weight (Fig. 3C) and 28.3%, 36.9%, and 41.1% lower dry weight (Fig. 3D) than the transgenic lines OE-4, OE-5, and OE-7, respectively. Similarly, the WT and transgenic plants exposed to drought stress exhibited lower fresh weight and dry weight than the untreated control. Nevertheless, the OE-4, OE-5, and OE-7 transgenic lines had 30.1%, 37.2%, and 42.6% more fresh weight than the WT, and 40.6%, 50.6%, and 60.4% more dry weight. These findings suggest that SikCOR413PM1 overexpression in cotton seedlings enhances the cold and drought tolerance.

Fig. 3
figure 3

Phenotypic analysis of 70-day-old WT and SikCOR413PM1-overexpressing cotton Seedlings under cold and drought stress. (A) Phenotype of WT and OE-4, OE-5, and OE-7 transgenic plants after exposure to 4 °C for 24 h and 48 h; (B) Phenotype of WT and OE-4, OE-5, and OE-7 transgenic plants after exposure to drought for 14 and 21 days; (C) Fresh weight and (D) dry weight of the plants under stress. The data shown are the average of three independent biological replicates. Bars represent standard deviations (SDs). * and ** indicate significant differences relative to WT cotton plants at P < 0.05 and P < 0.01, respectively (In the phenotypic analysis figure of cotton cold stress, the cotton seedlings in each figure are wild type, OE-4, OE-5, and OE-7 transgenic plants from left to right)

Overexpression of SikCOR413PM1 alleviated the damage to the cell membrane and enhanced the accumulation of osmoregulatory substances

To further elucidate the possible physiological mechanisms through which SikCOR413PM1 overexpression improves cell protection under cold and drought stress in cotton, we assessed the MDA content and relative conductivity of WT and transgenic plants (OE-4, OE-5, and OE-7). The MDA content in the leaves of transgenic plants was significantly lower than the WT plants under both stress conditions (Fig. 4A). Under 4 °C stress, the MDA content of OE-4, OE-5, and OE-7 was 31.7%, 31.6%, and 36.9% lower than that of the WT cotton. Under drought stress, the MDA content of OE-4, OE-5, and OE-7 was 26.3%, 27.5%, and 35.8% lower than that of WT cotton, respectively (Fig. 4B). Under normal conditions, the relative electrolyte leakage of both the WT and the transgenic lines was around 20%. However, the relative electrolyte leakage (REL) increased by 49.9%, 36.3%, 33.9%, and 39.4% after exposure to 4 °C and 46.6%, 33.2%, 29.5%, and 36.8% after drought treatment in the WT, OE-4, OE-5, and OE-7 plants, respectively.

Furthermore, the present study found low proline and soluble sugar levels in both WT and SikCOR413PM1-overexpressing plants before exposure to stress, and no significant difference between them (Fig. 4C and D). However, the levels of these two osmolytes significantly increased after exposure to cold and drought stress in WT and transgenic plants (OE-4, OE-5, and OE-7). Under 4 °C stress, OE-4, OE-5, and OE-7 exhibited 43.1%, 40.8%, and 55.1% higher proline levels and 32.0%, 38.6%, and 42.9% higher soluble sugar levels than the WT. Similarly, after drought treatment, OE-4, OE-5, and OE-7 had 41.8%, 42.7%, and 47.2% higher proline levels and 36.9%, 45.8%, and 47.9% higher soluble sugar levels than the WT. These observations indicate that SikCOR413PM1 overexpression maintains the integrity of the cell membrane and promotes the accumulation of osmotic protective agents in cotton under stress.

Fig. 4
figure 4

Physiological analysis of WT and SikCOR413PM1-overexpressing cotton plants after cold and drought stress treatments. (A) Malondialdehyde (MDA) content; (B) Relative electrolyte leakage (REL); (C) Proline content; (D) Soluble sugar content. The data shown for the WT are the means of three replicates, whereas those for the transgenic plants are means of three different lines. Bars represent standard deviations (SDs), * and ** indicate significant differences relative to WT cotton plants at P < 0.05 and P < 0.01

Overexpression of SikCOR413PM1 maintained high antioxidant enzyme activity and gene expression

The study further analyzed the O2− and H2O2 content in the WT and transgenic lines to assess their stress response. No significant difference was observed among the lines under normal conditions; however, after exposure to cold and drought stress, significant differences were detected between them. The O2− and H2O2 levels in the transgenic lines were significantly reduced compared to those in the WT under stress (Fig. 5A and B).

We further measured the activity of the antioxidant enzymes superoxide dismutase (SOD), peroxidase (POD), catalase (CAT), and glutathione S-transferase (GST). Under normal conditions, no significant difference was observed in enzyme activity between the transgenic lines and the WT. Exposure to stress enhanced the antioxidant enzyme activity in both plants. The transgenic lines exhibited higher antioxidant enzyme activities than the WT lines. Consistently, the expression levels of GhSOD, GhPOD, GhCAT, and GhGST in the transgenic lines were significantly higher than those in the WT lines after stress treatment (Fig. 5C–J).These observations indicate that SikCOR413PM1 overexpression increased the expression of genes encoding the antioxidant enzymes and enhanced the activity of the corresponding enzymes to regulate the amount of ROS and minimize the effect of oxidative stress.

Fig. 5
figure 5

Reactive oxygen species (ROS), antioxidant enzyme activity, and antioxidant gene expression in WT and SikCOR413PM1-overexpressing cotton plants under cold and drought stress. (A) O2 content; (B) H2O2 content; (C) SOD activity; (D) GhSOD expression; (E) POD activity; (F) GhPOD expression; (G) CAT activity; (H) GhCAT expression; (I) GST activity; (J) GhGST expression. The data shown for the WT are means of three replicates, whereas those for the transgenic plants are means of three different lines. Bars represent standard deviations (SDs). * and ** indicate significant differences relative to WT cotton plants at P < 0.05 and P < 0.01. The expression leve at 0h (control time point) was defined as 1.0

Analysis of stress-related gene expression in SikCOR413PM1-overexpressing cotton plants

We subsequently analyzed the expression levels of several stress-related genes, such as GhDREB1A, GhDREB1B, GhDREB1C, GhERF2, GhNAC3, and GhRD22, in cotton plants. The analysis revealed no significant difference in the expression levels of GhDREB1A, GhDREB1B, and GhDREB1C between the WT and transgenic lines under normal conditions. However, after exposure to cold and drought stress, the expression levels of these genes in transgenic lines were significantly higher than those in the WT lines. Under cold stress, the expression levels of GhDREB1A, GhDREB1B, and GhDREB1C genes in transgenic lines were 16.7, 9.9, and 8.6 times (OE-4), 16.7, 9.5, and 9.1 times (OE-5), and 18.7, 13.1, and 10.2 times (OE-7) higher than those in the WT, respectively. The transgenic lines showed different expression levels of these genes after drought stress treatment. Compared with the wild type, the expression levels in OE-4 lines were increased by 9.6, 17.7, and 5.5-fold, respectively. The corresponding expression folds for OE-5 and OE-7 lines were 9.1, 19.3, 4.9, and 11.0, 21.7, and 6.6, respectively. (Fig. 6A–C). Under cold stress, no significant difference was observed in GhRD22 expression between the transgenic and WT cotton. Under drought stress, the expression of GhRD22 was significantly upregulated and was 11.2-fold (OE-4), 13.1-fold (OE-5), and 15.1-fold (OE-7) higher than that of the WT, respectively (Fig. 6D). Additionally, GhERF2 expression in WT and transgenic plants was significantly different under stress treatment. The expression of GhERF2 in transgenic plants was 9.1 and 11.7 times (OE-4), 8.1 and 13.4 times (OE-5), and 13.8 and 15.9 times (OE-7) higher than that in the WT under cold and drought stress, respectively (Fig. 6E). Under drought stress, no significant difference was observed in GhNAC3 expression between transgenic and WT cotton. Meanwhile, after cold stress treatment, the expression levels increased in each plant. The expression levels of transgenic lines OE-4, OE-5, and OE-7 were 10.7, 12.9, and 14.5 times higher than those of the WT (Fig. 6F). Under cold stress, the transgenic lines exhibited significantly higher expression levels of all stress-related genes, except GhRD22. Similarly, under drought stress, the expression levels of all stress-related genes, except GhNAC3, were significantly higher in the transgenics than the WT. Besides, among the three transgenic lines, OE-7 exhbitied the highest expressionof these stress-related genes. These results indicate that SikCOR413PM1 overexpression improves the expression of stress response genes and enhances the cold and drought tolerance of cotton.

Fig. 6
figure 6

Relative expression of stress-related genes in WT and SikCOR413PM1-overexpressing cotton lines under cold and drought stress. Expression levels of (A) GhDREB1A; (B) GhDREB1B; (C) GhDREB1C; (D) GhRD22; (E) GhERF2; (F) GhNAC3, data shown are means ± SD of three replicates * and ** indicate significant differences relative to WT cotton plants at P < 0.05 and P < 0.01. The expression leve at 0h (control time point) was defined as 1.0

Agronomic traits of SikCOR413PM1-overexpressing cotton in the field

In the greenhouse, the plant height of transgenic cotton overexpressing SikCOR413PM1 increased slightly but not significantly, and the number of fruit branches per plant and bolls per plant increased significantly compared with the WT (Table. S1). The SikCOR413PM1-overexpressing lines had eight to nine fruit branches per plant, while the WT had only seven. In addition, the transgenic lines had an average of one or two bolls per plant, while the WT had. Although no significant difference was observed in boll weight and lint weight between the transgenic lines and WT plants, the fiber yield of the transgenic lines was moderately higher than that of the WT. The seed cotton yield and lint yield of transgenic cotton were also higher than those of the WT. These differences probably resulted from the SikCOR413PM1-mediated multiple tolerance activities.

Further, we performed field trials to evaluate the drought tolerance of cotton plants overexpressing SikCOR413PM1 (Fig. 7). The plant height, fruit branch number per plant, boll number per plant, and seed cotton yield of OE-4 and OE-7 lines were significantly higher than those of the WT. However, no significant difference was observed in the single boll weight, single boll lint weight, and lint percentage between the OE-4 and OE-7 lines. These results indicate that SikCOR413PM1 overexpression could improve the tolerance and yield of cotton under drought stress in the field.

Fig. 7
figure 7

Phenotypic analysis of WT and SikCOR413PM1-overexpressing cotton lines under drought stress in the field. (A) Plants grown under normal conditions with regular irrigation. Photos were taken on 70 days after drought stress. (B) Plants grown under drought stress. Photos were taken on 70 days after drought stress

Methods

Plant materials and growth conditions

The Xinjiang upland cotton (Gossypium hirsutum) Xinluzao 17 was used in this study. The seeds of this cotton were provided by Li Baocheng, the vice president of the Xinjiang Academy of Agriculture and Reclamation Sciences. Under normal conditions, the WT and transgenic plants were maintained in a tissue culture laboratory under 60–70% relative humidity, 16 h/8 h light/dark cycle, and 25 °C average temperature. The seedlings were grown in plastic pots containing a composite matrix of nutrient soil, vermiculite, and perlite (1:1:1) and maintained in a culture chamber with a light intensity of 500 µmol− 2 s− 1 at a temperature range of 22–28 °C and relative humidity of 60–70%.

Real-time quantitative PCR

Total RNA was isolated from the plant materials using the RNAiso Plus kit (TaKaRa) and treated with DNase I, following the manufacturer’s instructions. The first-strand cDNA was synthesized from the total RNA using oligo(dT) primers and PrimeScript® RTase (TaKaRa). Then, quantitative RT-PCR was performed on a LightCycler 480 instrument (Roche Diagnostics, Switzerland) with SYBR® Premix Ex TaqTM (TaKaRa, China) following the supplier’s instructions and MIQE guidelines to determine the expression patterns of SikCOR413PM1 and various stress-related genes. Each reaction mixture (10 µL) contained 50 ng cDNA, 5 µL 2× SYBR® Green MasterMix reagent (Applied Biosystems), and 0.2 µM gene-specific primers (Table.S2). The qRT-PCR program involved pre-denaturation at 95 °C for 60 s, followed by 40 amplification cycles with denaturation at 95 °C for 15 s, annealing at 61 °C for 30 s, and an extension at 72 °C for 30 s, and pre-denaturation at 52 °C for 40 s. GhHis3 (AF024716) was used as the reference gene, and the relative gene expression was calculated following the ΔΔCt method [63]. Each analysis was repeated three times.

Transgenic cotton generation: a comprehensive overview of the process

Our previous work identified and sequenced the SikCOR413PM1 gene. The WT plant without the exogenous gene were confirmed through molecular detection before the transformation, and the transgenic cotton plants containing SikCOR413PM1 were obtained using the pollen tube pathway-mediated method in this study. The WT plants at the flowering stage were transformed, and the seed cotton from the bolls was harvested and labeled. These T0 seeds were planted in isolated and protected experimental fields during the second year, and the seedlings with expanded true leaves were screened for kanamycin (2 g/L) resistance. Further, DNA was extracted from the putative ones for detecting the transgene by PCR. From the positive transgenics, total RNA was extracted for qRT-PCR identification. The positive transgenic plants screened and identified were routinely managed and grown to blossom and produce bolls. Seeds were collected from these transgenics and sown on 1/2 MS medium containing kanamycin (80 µg/mL) for further screening. The resistant seedlings were transferred to soil for growth, and the seeds were collected at maturity for additional screening up to T4 generation.

Plant stress treatments

Twenty-day-old WT and SikCOR413PM1-overexpressing transgenic cotton seedlings were subjected to 4 °C treatment for 24, 48, and 72 h and 20% PEG6000 for 10, 15, and 20 days. The phenotype of the cold-treated plants and the phenotype, the survival rate, fresh weight, and dry weight of the drought-treated plants were analyzed at regular intervals. Similarly, 70-day-old WT and SikCOR413PM1-overexpressing transgenic cotton plants were exposed to 4 °C for 24 and 48 h and 20% PEG6000 for 14 and 21 days. After both stress treatments, the plant phenotypic changes were recorded, and the leaves were collected from similar positions on the plants. And the collected samples were used for subsequent analysis of the fresh and dry weights, physiological indexes, enzyme activities, and gene expression levels.

Investigation of agronomic traits

The plant height of selected WT and transgenic plants were measured using a tape, and the stem diameter using an electronic vernier caliper. The plant height was measured from the root neck to the apex. The stem diameter was measured 1 cm above the first leaf. The fresh weight and dry weight of the leaves. were measured using an electronic weighing scale.

Analysis of the agronomic traits of cotton under drought stress in the field

Independent T0 transgenic plants were grown until maturity in the field. The inbred line seeds (T1) were harvested and sown in the experimental field. Further, based on the expression of the kanamycin resistance gene, T2 seeds were screened from the positive T1 lines. The homozygous transgenic lines were selected for selfing to obtain the T3 generation. This selection was carried out until the T4 generation to obtain a stable transgenic line. The performance of T4 homozygous transgenic and WT cotton lines was evaluated during the growing season in normal and drought-prone areas.

The traits of the WT and transgenic plants under normal environmental conditions were analyzed in the experimental field of the Agricultural Biotechnology Key Laboratory of Shihezi University in Shihezi City, Xinjiang. After one month, the plants were exposed to drought stress. All field trials were carried out using a randomized complete block design with three replicates. Each transgenic cotton was planted in a test plot of about 10.25 m2, with 240 plants per field, and repeated three timesice. A plant spacing of 12.5 cm and a row spacing of 30 cm were adopted. All recommended plant protection measures were carried out from sowing to harvest. Further, 50 plants were randomly selected from each replicate for analyzing the agronomic traits, such as plant height, fruit branch number per plant, boll number per plant, boll weight, lint weight per boll, seed cotton yield per plant, lint yield per plant, lint percentage, seed cotton yield, and lint yield per plot. The height of the main stem at the boll opening stage was represented as the plant height. The vegetative branches and the fruit branches were manually separated from the main stem and counted. Bolls were counted to determine the bolls per plant. Besides, all naturally opened bolls were collected from a plot and dried in an oven at 37 °C. Then, the weight of 100 randomly selected the boll weight, lint cotton yield, unginned cotton yield and lint percentage was measured.

Statistical analysis

GraphPad Prism (version 7.0; GraphPad Software, LLC225 Franklin Street. Fl. 26 BOSTON, MA 02110 USA) and SPSS software (version 11.0; SPSS, Chicago, IL, USA) were used for statistical analysis. The data were expressed as the mean ± standard deviation of three independent experiments. The graphs were plotted using CorelDRAW X4 SP2 software (version 14.0; CorelDRAW X4 SP2, Canada). Dunnett’s multiple comparison test assessed the differences between WT and transgenic lines. The differences between the transgenic lines and WT plants were considered statistically significant at P < 0.05 and P < 0.01.

Discussion

Abiotic stresses are known to cause water deficit in plants [32, 64, 65]. However, plants have evolved physiological mechanisms to adapt to water shortage, such as accumulating various osmotic regulators and antioxidants and producing ROS [32, 66]. Under cold and drought conditions, they maintain cell water by accumulating osmotic solutes, regulating osmotic potential, and adjusting osmotic pressure [67, 68]. They also initiate defense mechanisms by increasing the concentration of sugars, proline, amino acids, and other substances, reducing cell osmotic potential and increasing water absorption [69].

In this study, the soluble protein and proline levels in the SikCOR413PM1-overexpressing lines were significantly higher than those in the WT under cold and drought stress (Fig. 4C and D). Generally, the stronger the osmotic adjustment ability, the higher the tolerance and adaptability to osmotic stress. Thus, these observations suggest that SikCOR413PM1 overexpression enhances the environmental adaptability of cotton by improving the osmotic adjustment ability through the accumulation of osmotic adjustment substances.

Typically, any change in environmental conditions affects cell membrane function [70,71,72]. Nevertheless, plants stabilize cell membranes under cold stress to prevent freezing-induced dehydration [73]. Some studies have shown that soluble sugars and proline stabilize proteins and overall cell structure, enhance cell membrane integrity, and improve frost resistance [74,75,76,77,78]. In cotton, the soluble sugars and proline protect cell membranes from oxidative damage under various abiotic stresses [79]. Proline, the most common osmotic solute, prevents water loss, reduces mechanical damage, and enhances stress tolerance [75, 80, 81]. In the present study, SikCOR413PM1 overexpression in cotton enhanced the accumulation of soluble sugars and proline under cold and drought environments (Fig. 4C–D), which probably improved cell membrane stability.

The cell membrane is the primary target of adverse stress, and maintaining membrane integrity and stability under abiotic stress is critical for tolerance [82]. Under cold stress, the structure and function of plant cell membrane change, and membrane lipid peroxidation occurs, increasing cell membrane permeability, ROS, and relative conductivity [83]. Thus, the content of MDA reflects the degree of membrane lipid peroxidation and cell damage [84], and relative conductivity indicates plasma membrane damage caused by freezing [83].

There were no significant differences between WT and transgenic plants under normal conditions compared to REL and MDA. However, under 4 °C and drought stress, the contents of MDA and REL increased to different degrees, and the growth of WT was higher than that of transgenic plants. MDA and REL reduce the degree of plasma membrane damage by increasing their own content, thereby improving the stress resistance of cotton. The experimental results showed that MDA and REL increased more in WT than in transgenic plant, which indicated that transgenic plant was more tolerant to cold and drought stress than WT. Plants can also enhance freezing tolerance by improving antioxidant mechanisms [65, 85]. Thus, the lower content of MDA in SikCOR413PM1-overexpressing cotton than that of the WT indicates reduced membrane lipid peroxidation and enhanced stress resistance.

Several studies have investigated whether COR genes are induced in response to water shortage and cold stress. In Arabidopsis, water shortage induced most COR genes [86,87,88]. Similarly, cold and drought stresses induced COR25 to protect Brassica napus cells from dehydration [35]. Therefore, we speculate that drought and cold stresses induce SikCOR413PM1 in cotton, and the upregulation protects cell membranes and reduces cell damage by regulating osmotic regulators, MDA, and relative conductivity.

Currently, research on COR413 family genes is limited. However, Zhang et al. found that the overexpression of LeCOR413PM2 located in the plasma membrane enhanced the tolerance of transgenic tomatoes to cold stress by protecting the membrane from cold damage [45]. Similarly, the overexpression of SikCOR413PM1, located on the plasma membrane, improved tobacco’s drought and cold tolerance by protecting cell membrane stability [43]. It has been predicted that COR413 family proteins located in the plasma membrane (COR413-PM) have a film-forming function. These earlier reports suggested that COR413 family proteins in the plasma membrane play significant roles in maintaining cell membrane stability, requiring further investigation.

Under extreme temperatures and water deficit conditions, plants experience cell dehydration, photosynthetic damage, and membrane oxidation [32]. The abiotic stresses interfere with photosynthetic electron transport and produce ROS, severely damaging the chloroplasts [67, 89,90,91]. The organelles with high oxidative activity, such as mitochondria, chloroplasts, and peroxisomes, act as the primary sources of ROS and generate O2 and H2O2, leading to oxidative damage of proteins, DNA, and lipids [16, 18, 90, 92]. Therefore, regulating ROS can protect cells from oxidative damage [93, 94].

ROS signaling is an adaptive regulatory mechanism of plants to stressful environments [67]. Plants have established an efficient ROS scavenging system, including antioxidant enzymes and antioxidants, which reduce ROS-induced damage [95]. Cell-protective enzymes mainly include SOD, CAT, APX, and GR. Overexpression of genes encoding these enzymes improves plant ability to scavenge ROS and their tolerance to abiotic stress [96,97,98,99].

The study also detected O2 and H2O2 in WT and transgenic lines under cold and drought stress. The transgenics had lower O2 and H2O2 content than the WTunder cold and drought stress (Fig. 5A-B). These observations indicate that SikCOR413PM1 overexpression enhances the plant’s ability to scavenge ROS. Moreover, all stress treatments enhanced the activity of antioxidant enzymes in plants, particularly the transgenic lines. The expression levels of GhSOD, GhPOD, GhCAT, and GhGST in transgenic lines were also significantly higher than those in the WT lines after stress treatment (Fig. 5C–J). These results collectively suggest that the overexpression of SikCOR413PM1 increased the expression level of genes corresponding to antioxidant enzymes in transgenic cotton, thereby enhancing the activity of enzymes that regulate ROS to reduce oxidative stress. Therefore, the study concludes that SikCOR413PM1 maintains high antioxidant enzyme activity and gene expression, reducing oxidative stress in cotton.

Previous studies have related cold-induced changes in antioxidant enzyme activity in the leaves with frost resistance in several cereals [100,101,102]. Overexpression of the SOD-encoding gene in tobacco and the APX-encoding gene in tomato increased their tolerance to cold [103, 104]. The antioxidant defense system is closely related to plant frost resistance [90, 95, 105]. Cold acclimation induces ROS accumulation, activating the antioxidant defense system [106]. Mariem Ben Abdallah et al. indicate that in olive, drought stress pre-exposure can maintain homeostasis by increasing the activity of ROS scavenge enzymes (such as guaiacol peroxidase, SOD, and CAT) and reducing H2O2 and MDA levels [107].Chen et al. (2015) [108] found that silencing GbMYb5 resulted in higher oxidative stress in cotton under drought conditions due to decreased POD, SOD, and CAT activities. These reports further confirm that overexpression of SikCOR413PM1 can improve the cold and drought tolerance of cotton by reducing oxidative stress.

Proline protects cells from increased stress-related ROS levels [67]. In a previous study, P5CS transgenic wheat accumulated higher proline than WT plants and showed lesser membrane lipid peroxidation during drought [109]. This study indicated that proline plays a role in reducing ROS damage during drought. Specifically, cold acclimation induces rapid proline accumulation by changing the activity of proline metabolic pathway-related enzymes, thereby improving frost resistance [110]. In this study, transgenic cotton used proline as an osmotic regulator to improve cold and drought tolerance. The proline content was higher in the transgenic plants under drought and cold stresses than in the WT plants. Thus, we speculate that SikCOR413PM1 overexpression increased proline accumulation, reducing the impact of ROS on cells.

Furthermore, the DREB family of transcription factors has been shown to play an important role in cold adaptation-induced frost tolerance. These transcription factors induce the expression of cold and dehydration-regulated genes in plants [111,112,113]. Studies have reported enhanced expression of DREB genes under cold and drought conditions [114]. For example, cold stress induced DREB1A, DREB1B, and DREB1C expression in Arabidopsis [113]. Thus, the increase in the expression of DREB partly explains the resistance of plants to stress.

In this study, the expression levels of GhDREB1A, GhDREB1B, and GhDREB1C were significantly higher in transgenic lines than in WT lines after cold and drought stress treatment (Fig. 6A–C). The expression levels of GhRD22, GhERF2, and GhNAC3 also increased under both stresses (Fig. 6D–F). These results indicate that drought and cold strongly induced GhDREB1A, GhDREB1B, and GhDREB1C, and the expression of the five stress-related genes in the transgenic cotton plants were higher than in WT. These observations suggest that SikCOR413PM1 overexpression helps increase the expression level of stress response genes. In other words, the increase in the expression of DREB1 in the transgenics indicates a stronger tolerance than the WT under cold and drought conditions. In Arabidopsis, the DREB1 protein activates COR genes, such as COR15a [115], COR6.6 [116], and COR47 [117], and DREB1 protein plays a functional role in enhancing tolerance to cold and drought stress. Therefore, it can be speculated that stress increases the expression of DREB1, which may be related to the activation of the COR413 gene. In summary, the overexpression of SikCOR413PM1 promoted the expression level of stress genes, such as DREB1, thereby improving resistance to cold and drought stresses.

Overexpression of SikCOR413PM1 positively influenced the growth of cotton under greenhouse conditions. Then, transgenic plants’ performance was studied in the field to investigate its actual impact on cotton production. The transgenic cotton had significantly higher seed and lint yields in the field, indicating its resistance to yield loss under stress. Although the growth status of the two lines did not differ significantly under normal conditions, the OE-4 and OE-7 transgenic lines exhibited substantially better growth under drought conditions than the WT (Fig. 7). Moreover, the number of fruit branches per plant, number of bolls per plant, and seed cotton yield were also significantly higher in the transgenic cotton than the WT (Table. S1). These results suggest that SikCOR413PM1 overexpression improved drought tolerance and yield of cotton. The findings also provide a theoretical foundation for developing cotton varieties with improved environmental adaptability and yield.

In nature, plants often face multiple abiotic stresses, significantly impacting growth and yield. So, to enhance cold and drought tolerance, the SikCOR413PM1 gene was overexpressed in cotton plants. For stress analysis, three transgenic lines with high expression levels were selected (OE-4/OE-5/OE-7). The results demonstrated that overexpression of SikCOR413PM1 improved cotton’s ability to tolerate cold and drought stress through various mechanisms. Overexpression of SikCOR413PM1 can accumulate osmotic substances by enhancing osmotic adjustment ability, improve ROS scavenging mechanism by increasing antioxidant enzyme activity and related gene expression, and maintain the integrity and stability of cell membrane by up-regulating stress-related genes, and finally alleviate the yellowing, withering, falling and waterlogging of leaves of transgenic lines in the face of low temperature and drought stress. Because the cell membrane of transgenic lines is more complete and stable, some plants can resume growth under normal conditions after stress. These findings indicate SikCOR413PM1 is a candidate for breeding cotton varieties with improved cold and drought tolerance (Fig. 8). This research advances our knowledge and comprehensively explains how SikCOR413PM1 enhances cotton’s resilience against cold and drought stresses. However, further studies are required to understand better the molecular mechanisms underlying SikCOR413PM1’s ability to mediate tolerance to cold and drought stress. These studies should focus on unraveling the signaling network and the specific functions of SikCOR413PM1.

Fig. 8
figure 8

Illustration of the Mechanism by which the SikCOR413PM1 Gene Mitigates Cold and Drought Stress

The SikCOR413PM1 gene enhances cotton’s overall stress tolerance by upregulating stress-related genes and antioxidant enzymes. Furthermore, SikCOR413PM1 enhances cell osmotic regulation capabilities, while concurrently reducing cell membrane oxidative damage and membrane lipid peroxidation. This is achieved through increased antioxidant enzyme activity, elevated soluble substance levels, and reduced malondialdehyde content. Ultimately, these mechanisms lead to improved plant tolerance and increased crop yields.

Data Availability

We submitted the gene sequence at NCBI, and since the gene is not publicly available, only the GenBank accession number of the nucleotide sequence provided by NCBI is available: BankIt2706231 Seq OR037196.

Other GenBank accession numbers involved in the article are as follows:

GhDREB1A:GenBank: AY321150.3.

GhDREB1B: Ghi_A04G04116.

GhDREB1C:GenBank: AYL99847.1.

GhNAC3:GenBank: ACI15349.1.

GhRD22:GeneID: 107,959,794.

GhERF2:GenBank: DQ464375.1.

GhHis3: GenBank: AF024716.

Abbreviations

ABA:

Abscisic acid

MET:

Melatonin

BRs:

Brassinosteroids

CAT:

Catalase

CBF:

C-receptor binding factor

COR:

Cold-regulated

CRT:

C-receptor

DREB:

Dehydration response element binding factor

GST:

Glutathione S-transferase

POD:

Peroxidase

qRT-PCR:

Quantitative Real-Time PCR

REL:

Relative electrical conductivity

ROS:

Reactive oxygen species

SOD:

Superoxide Dismutase

MDA:

Malondialdehyde

FT:

Freezing tolerance

GR:

Glutathione reductase

REL:

Relative electrolyte leakage

References

  1. Varshney RK, Barmukh R, Roorkiwal MQi, Kholova J, Tuberosa R et al. Breeding custom-designed crops for improved drought adaptation. Adv Genet.2021; 2, e202100017.

  2. Mboup M, Fischer I, Lainer H, Stephan W. Trans-species polymorphism and allele-specific expression in the CBF gene family of wild tomatoes. Mol Biol Evol. 2012;29(12):3641–52.

    Article  PubMed  CAS  Google Scholar 

  3. Gerik TJ, Faver KL, Thaxton PM. Late season water stress in cotton: I. Plant Growth, Water Use, and yield. Crop Sci. 1996;36(4):914.

    Article  Google Scholar 

  4. Raza A, Mubarik MS, Sharif R, Habib M, Jabeen W, Zhang C, Chen H, Chen ZH, Siddique KHM, Zhuang W, Varshney RK. Developing drought-smart, ready-to-grow future crops. Plant Genome. 2023;16(1):e20279.

    Article  PubMed  Google Scholar 

  5. Nabwire S, Wakholi C, Faqeerzada MA. Estimation of cold stress, Plant Age, and number of leaves in Watermelon plants using image analysis. Front Plant Sci. 2022;13:847225.

    Article  PubMed  PubMed Central  Google Scholar 

  6. Albertos P, Wagner K, Poppenberger B. Cold stress signalling in female reproductive tissues. Plant Cell Environ. 2019;42(3):846–53.

    Article  PubMed  CAS  Google Scholar 

  7. Thakur P, Kumar S, Malik JA, et al. Cold stress effects on reproductive development in grain crops: an overview. Environ Exp Bot. 2010;67(3):429–43.

    Article  CAS  Google Scholar 

  8. Arshad MS et al. Thermal stress impacts reproductive development and grain yield in rice.Plant Physiol.Biochem. 2017.

  9. Zhang W, Wang J, Huang Z, et al. Effects of low temperature at Booting Stage on sucrose metabolism and endogenous hormone contents in Winter Wheat Spikelet. Front Plant Sci. 2019;10:498.

    Article  PubMed  PubMed Central  Google Scholar 

  10. Galeano E, Vasconcelos TS, Novais de Oliveira P, et al. Physiological and molecular responses to drought stress in teak (Tectona grandis L.f). PLoS ONE. 2019;14(9):e0221571.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Abid Ullah H, Sun X, Yang et al. Drought coping strategies in cotton: increased crop per drop. Plant Biotechnol J 2017.15(3): 271–84.

  12. Haddad N, Choukri H, Ghanem ME, et al. High-Temperature and Drought stress effects on Growth, Yield and Nutritional Quality with Transpiration Response to Vapor pressure deficit in Lentil. Plants (Basel). 2021;11(1):95.

    Article  PubMed  Google Scholar 

  13. Raza A, Charagh S, Salehi H, Abbas S, Saeed F, Poinern GEJ, et al. Nano-enabled stress-smart agriculture: can nanotechnology deliver drought and salinity-smart crops? J Sustain Agric Environ. 2023;2:189–214.

    Article  Google Scholar 

  14. Walter J, Nagy L, Hein R, Rascher U, Beierkuhnlein C, Willner E, Jentsch A. Do plants remember drought? Hints towards a drought-memory in grasses. Environ Exp Bot. 2011;71:34–40.

    Article  Google Scholar 

  15. Hossain MA, Burritt DJ, Fujita M. Cross-stress tolerance in plants: molecular mechanisms and possible involvement of reactive oxygen species and methylglyoxal detoxification systems. In: Tuteja N, Gill SS, editors Abiotic stress response in plants. 2016;323–375.

  16. Sharma P, Jha AB, Dubey RS, Pessarakli M. Reactive oxygen species, oxidative damage, and antioxidative defense mechanism in plants under stressful conditions. J Bot. 2012;2012:217037.

    Google Scholar 

  17. Guler NS, Pehlivan N. Exogenous low-dose hydrogen peroxide enhances drought tolerance of soybean (Glycine max L.) through inducing antioxidant system. Acta Biol Hung. 2016;67(2):169–83.

    Article  PubMed  CAS  Google Scholar 

  18. Tripathy BC, Oelmüller R. Reactive oxygen species generation and signaling in plants. Plant Signal Behav. 2012;7(12):1621–33.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  19. Bhattacharjee S. Reactive oxygen species and oxidative burst: roles in stress, senescence and signal transduction in plants. Curr Sci. 2005;89:1113–21.

    CAS  Google Scholar 

  20. Baier M, Bittner A, Prescher A, Buer, Jvan. Preparing plants for improved cold tolerance by priming. Plant Cell Environ. 2019;42:782–800.

    Article  PubMed  CAS  Google Scholar 

  21. Chinnusamy V, Zhu J, Zhu JK. Cold stress regulation of gene expression in plants. Trends Plant Sci. 2007;12:444–51.

    Article  PubMed  CAS  Google Scholar 

  22. Lv B, Tian H, Zhang F, Liu J, Lu S, Bai M, Li C, Ding Z. Brassinosteroids regulate root growth by controlling reactive oxygen species homeostasis and dual effect on ethylene synthesis in Arabidopsis. PLoS Genet. 2018;14, e1007144.

  23. Fang P, Yan M, Chi C, Wang M, Zhou Y, Zhou J, Shi K, Xia X, Foyer CH, Yu J. Brassinosteroids act as a positive regulator of photoprotection in response to chilling stress. Plant Physiol. 2019;180:2061–76.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  24. Raza A, Charagh S, García-Caparrós P, Rahman MA, Ogwugwa VH, Saeed F, Jin W. Melatonin-mediated temperature stress tolerance in plants. GM Crops Food. 2022;13(1):196–217.

    Article  PubMed  PubMed Central  Google Scholar 

  25. Raza A, Charagh S, Najafi Kakavand S, Abbas S, Shoaib, Yasira, Anwar, Sultana, Sharifi, Sara, Lu, Guangyuan, Siddique, Kadambot. Role of phytohormones in regulating cold stress tolerance. Physiological and molecular approaches for developing cold-smart crop plants; 2023.

  26. Ning J, Li X, Hicks LM, Xiong L. A raf-like MAPKKK gene DSM1 mediates drought resistance through reactive oxygen species scavenging in rice. Plant Physiol. 2010;152:876–90.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. You J, Zong W, Hu H, Li X, Xiao J, Xiong L. A STRESS-RESPONSIVE NAC1-regulated protein phosphatase gene rice protein phosphatase18 modulates drought and oxidative stress tolerance through abscisic acid-independent reactive oxygen species scavenging in rice. Plant Physiol. 2014;166:2100–14.

    Article  PubMed  PubMed Central  Google Scholar 

  28. Hazman M, Hause B, Eiche E, Nick P, Riemann M. Increased tolerance to salt stress in OPDA-deficient rice ALLENE OXIDE CYCLASE mutants is linked to an increased ROS-scavenging activity. J Exp Bot. 2015;66:3339–52.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  29. Zhao C, Lang Z, Zhu J-K. Cold responsive gene transcription becomes more complex. Trends Plant Sci. 2015;20:466–8.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  30. Ding Y, Shi Y, Yang S. Advances and challenges in uncovering cold tolerance regulatory mechanisms in plants. New Phytol. 2019;222:1690–704.

    Article  PubMed  Google Scholar 

  31. Zhu J, Dong CH, Zhu JK. Interplay between cold-responsive gene regulation, metabolism and RNA processing during plant cold acclimation. Curr Opin Plant Biol. 2007;10(3):290–5.

    Article  PubMed  CAS  Google Scholar 

  32. Zhu JK. Abiotic stress signaling and responses in plants. Cell. 2016;167:313–24.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  33. Liu J, Shi Y, Yang S. Insights into the regulation of C-repeat binding factors in plant cold signaling. J Integr Plant Biol. 2018;60(9):780–95.

    Article  PubMed  Google Scholar 

  34. Guo X, Liu D, Chong K. Cold signaling in plants: insights into mechanisms and regulation. J Integr Plant Biol. 2018;60(9):745–56.

    Article  PubMed  Google Scholar 

  35. Chen L, Zhong H, Ren F, Guo QQ, Hu XP, Li XB. A novel cold-regulated gene, COR25, of Brassica napus is involved in plant response and tolerance to cold stress. Plant Cell Rep. 2011;30(4):463–71.

    Article  PubMed  Google Scholar 

  36. Wathugala DL, Richards SA, Knight H, Knight MR. OsSFR6 is a functional rice orthologue of SENSITIVE TO FREEZING-6 and can act as a regulator of COR gene expression, osmotic stress and freezing tolerance in Arabidopsis. New Phytol. 2011;191(4):984–95.

    Article  PubMed  CAS  Google Scholar 

  37. Wan F, Pan Y, Li J, Chen X, Pan Y, Wang Y, Tian S, Zhang X. Heterologous expression of Arabidopsis C-repeat binding factor 3 (AtCBF3) and cold-regulated 15A (AtCOR15A) enhanced chilling tolerance in transgenic eggplant (Solanum melongena L). Plant Cell Rep. 2014;33(12):1951–61.

    Article  PubMed  CAS  Google Scholar 

  38. Yuan HM, Sheng Y, Chen WJ, Lu YQ, Tang X, Ou-Yang M, Huang X. Overexpression of Hevea brasiliensis HbICE1 enhances Cold Tolerance in Arabidopsis. Front Plant Sci. 2017;8:1462.

    Article  PubMed  PubMed Central  Google Scholar 

  39. Breton G, Danyluk J, Charron JB, Sarhan F. Expression profiling and bioinformatic analyses of a novel stress-regulated multispanning transmembrane protein family from cereals and Arabidopsis. Plant Physiol. 2003;132(1):64–74.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  40. Okawa K, Nakayama K, Kakizaki T, Yamashita T, Inaba T. Identification and characterization of Cor413im proteins as novel components of the chloroplast inner envelope. Plant cell and Environment. 2008;31(10):1470–83.

    Article  CAS  Google Scholar 

  41. Zhang C, Li C, Liu J, Lv Y, Yu C, Li H, Zhao T, Liu B. The OsABF1 transcription factor improves drought tolerance by activating the transcription of COR413-TM1 in rice. J Exp Bot. 2017;68(16):4695–707.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Ma X, Wang G, Zhao W, Yang M, Ma N, Kong F, et al. SlCOR413IM1, a novel cold regulation gene from tomato, enhances drought stress tolerance in Tobacco. J Plant Physiol. 2017;216:88–99.

    Article  PubMed  CAS  Google Scholar 

  43. Guo X, Zhang L, Dong G, Xu Z, Li G, Liu N, Wang A, Zhu J. A novel cold-regulated protein isolated from Saussurea involucrata confers cold and drought tolerance in transgenic Tobacco (Nicotiana tabacum). Plant Science: An International Journal of Experimental Plant Biology. 2019;289:110246.

    Article  PubMed  CAS  Google Scholar 

  44. Zhou A, Liu E, Li H, Li Y, Feng S, Gong S, Wang J. PsCor413pm2, a plasma Membrane-Localized, Cold-regulated protein from Phlox subulata, confers low temperature tolerance in Arabidopsis. Int J Mol Sci. 2018;19(9):2579.

    Article  PubMed  PubMed Central  Google Scholar 

  45. Zhang L, Guo X, Zhang Z, Wang A, Zhu J. Cold-regulated gene LeCOR413PM2 confers cold stress tolerance in tomato plants. Gene. 2021;764:145097.

    Article  PubMed  CAS  Google Scholar 

  46. Wendel JF, Grover CE. Taxonomy and evolution of the cotton genes, Gossypium. In: Fang DD, Percy RG, editors. Cotton. Volume 2. Madison: ASA, CSSA, and SSSA; 2015. pp. 25–44.

    Chapter  Google Scholar 

  47. Lei W, Ben F. Analysis of the competitiveness of the cotton industry in Xinjiang based on the diamond model. UTMS J Econ. 2021;12:123–46.

    Google Scholar 

  48. Zheng J, Zhang Z, Liang Y, Gong Z, Zhang N, Ditta A, Sang Z, Wang J, Li X. Whole Transcriptome Sequencing Reveals Drought Resistance-Related Genes in Upland Cotton. Genes (Basel). 2022; 27;13(7):1159.

  49. Ullah A, Sun H, Yang X, Zhang X. Drought coping strategies in cotton: increased crop per drop. Plant Biotechnol J. 2017;15(3):271–84.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Cao Z, Zhao T, Wang L, Han J, Chen J, Hao Y, Guan X. The lincRNA XH123 is involved in cotton cold-stress regulation. Plant Mol Biol. 2021;106(6):521–31.

    Article  PubMed  CAS  Google Scholar 

  51. Mittler R. Abiotic stress, the field environment and stress combination. Trends Plant Sci. 2006;11(1):15–9.

    Article  PubMed  CAS  Google Scholar 

  52. Basso MF, Costa JA, Ribeiro TP, Arraes FBM, Lourenço-Tessutti IT, Macedo AF, Neves MRD, Nardeli SM, Arge LW, Perez CEA, Silva PLR, de Macedo LLP, Lisei-de-Sa ME, Santos Amorim RM, Pinto ERC, Silva MCM, Morgante CV, Floh EIS, Alves-Ferreira M. Grossi-De-sa MF. Overexpression of the CaHB12 transcription factor in cotton (Gossypium hirsutum) improves drought tolerance. Plant Physiol Biochem. 2021;165:80–93.

    Article  PubMed  CAS  Google Scholar 

  53. Liu J, Magwanga RO, Xu Y, Wei T, Kirungu JN, Zheng J, Hou Y, Wang Y, Agong SG, Okuto E, Wang K, Zhou Z, Cai X, Liu F. Functional characterization of cotton C-Repeat binding factor genes reveal their potential role in cold stress tolerance. Front Plant Sci. 2021;12:766130.

    Article  PubMed  PubMed Central  Google Scholar 

  54. Zhang B, Chang L, Sun W, Ullah A, Yang X. Overexpression of an expansin-like gene, GhEXLB2 enhanced drought tolerance in cotton. Plant Physiol Biochem. 2021;162:468–75.

    Article  PubMed  CAS  Google Scholar 

  55. Lu G, Wang L, Zhou L, Su X, Guo H, Cheng H. Overexpression of AmCBF1 enhances drought and cold stress tolerance, and improves photosynthesis in transgenic cotton. PeerJ. 2022;10:e13422.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Liu J, Mehari TG, Xu Y, Umer MJ, Hou Y, Wang Y, Peng R, Wang K, Cai X, Zhou Z, Liu F. GhGLK1 a key candidate gene from GARP Family enhances Cold and Drought stress tolerance in cotton. Front Plant Sci. 2021;12:759312.

    Article  PubMed  PubMed Central  Google Scholar 

  57. Tajo SM, Pan Z, Jia Y, He S, Chen B, Sadau SB, Km Y, Ajadi AA, Nazir MF, Auta U, Geng X, Du X. Silencing of GhORP_A02 enhances drought tolerance in Gossypium hirsutum. BMC Genomics. 2023 Jan;9 24(1):7.

  58. Chik WI, Zhu L, Fan LL, Yi T, Zhu GY, Gou XJ, Chen HB. Saussurea involucrata: a review of the botany, phytochemistry and ethnopharmacology of a rare traditional herbal medicine. J Ethnopharmacol. 2015;172:44–60.

    Article  PubMed  CAS  Google Scholar 

  59. Hua K, Zhang J, Botella JR, Ma C, Kong F, Liu B, Zhu JK. Perspectives on the application of Genome-Editing technologies in Crop breeding. Mol Plant. 2019;12(8):1047–59.

    Article  PubMed  CAS  Google Scholar 

  60. Liu B, Mu J, Zhu J et al. Saussurea involucrata SiDhn2 gene confers tolerance to drought stress in upland cotton. 2017.

  61. Sato H, Todaka D, Kudo M, Mizoi J, Kidokoro S, Zhao Y, Kazuko Yamaguchi-Shinozaki. The A rabidopsis transcriptional regulator DPB 3‐1 enhances heat stress tolerance without growth retardation in rice. Plant Biotechnol J. 2016;14(8):1756–67.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  62. Yu X, Brown JM, Graham SE, Carbajal EM, Zuleta MC, Milla-Lewis S. Detection of quantitative trait loci associated with drought tolerance in st. augustinegrass. PLoS ONE. 2019; 14(10).

  63. Livak KJ, Schmittgen TD. Analysis of relative gene expression data using real-time quantitative PCR and the 2 (T) (-Delta Delta C) method. Methods. 2001;25:402–8.

    Article  PubMed  CAS  Google Scholar 

  64. Steponkus PL. Role of the plasma membrane in freezing injury and cold acclimation. Annu Rev Plant Physiol. 1984;35:543–84.

    Article  CAS  Google Scholar 

  65. Thomashow MF. Plant cold acclimation: freezing tolerance genes and regulatory mechanisms. Annu Rev Plant Physiol Plant Mol Biol. 1999;50:571–99.

    Article  PubMed  CAS  Google Scholar 

  66. Xiong L, Schumaker KS, Zhu JK. Cell signaling during cold, drought, and salt stress. Plant Cell. 2002;14(Suppl):165–83.

    Article  Google Scholar 

  67. Miller G, Suzuki N, Ciftci-Yilmaz S, Mittler R. Reactive oxygen species homeostasis and signaling during drought and salinity stresses. Plant Cell Environ. 2010;33:453–67.

    Article  PubMed  CAS  Google Scholar 

  68. Mitra J. Genetics and genetic improvement of drought resistance in crop plants.Curr. Sci (Bangalore). 2001;80:758–63.

    CAS  Google Scholar 

  69. Fang Y, Xiong L. General mechanisms of drought response and their application in drought resistance improvement in plants. Cell Mol Life Sci. 2015;72(4):673–89.

    Article  PubMed  CAS  Google Scholar 

  70. Webb MS, Steponkus PL. Freeze-lnduced membrane ultrastructural alterations in rye (sede cereale) leaves. Plant Physiol. 1993;101:955–63.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  71. Uemura M, Joseph RA, Steponkus PL. Cold acclimation of Arabidopsis thaliana (effect on plasma membrane lipid composition and freeze-lnduced lesions). Plant Physiol. 1995;109:15–30.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  72. Ouellet F. Cold acclimation and freezing tolerance in plants. Encycl Life Sci. 2007.

  73. Steponkus PL, Webb MS. Freeze-induced dehydration and membrane destabilization in plants. In: Water and Life: Comparative Analysis of Water Relationships at the Organismic, Cellular and Molecular Level. 1992; 338–362.

  74. Guy CL. Cold acclimation and freezing stress tolerance: role of protein metabolism. Annu Rev Plant Physiol Plant Mol Biol. 1990;41:187–223.

    Article  CAS  Google Scholar 

  75. Ashraf M, Foolad MR. Roles of glycine betaine and proline in improving plant abiotic stress resistance. Environ Exp Bot. 2007;59:206–16.

    Article  CAS  Google Scholar 

  76. Valliyodan B, Nguyen HT. Understanding regulatory networks and engineering for enhanced drought tolerance in plants. Curr Opin Plant Biol. 2006;9:189–95.

    Article  PubMed  CAS  Google Scholar 

  77. Janska´ A, Marsı´k P, Zelenkova´ S, Ovesna´ J. Cold stress and acclimation—what is important for metabolic adjustment? Plant Biol. 2010;12:395–405.

    Article  PubMed  Google Scholar 

  78. Raza A, Tabassum J, Kudapa H, Varshney RK. Can omics deliver temperature resilient ready-to-grow crops? Crit Rev Biotechnol. 2021;41(8):1209–32.

    Article  PubMed  Google Scholar 

  79. Khan MS, Ahmed D, Khan MA. Utilization of genes encoding osmoprotectants in transgenic plants for enhanced abiotic stress tolerance. Electron J Biotechnol. 2015;18:257–66.

    Article  Google Scholar 

  80. Kovacs Z, Simon-Sarkadi L, Sovany C, Kirsch K, Galiba G, Kocsy G. Differential effects of cold acclimation and abscisic acid on free amino acid composition in wheat. Plant Sci. 2011;180:61–8.

    Article  PubMed  CAS  Google Scholar 

  81. Bartels D, Sunkar R. Drought and salt tolerance in plants.Critical Reviews in Plant Sciences. 2005; 24, 23–58.

  82. Levitt J. Responses of plants to Environmental stresses: water, radiation, salt and other stresses. Volume II. New York: Academic Press; 1980. pp. 3–211.

    Google Scholar 

  83. Jiang B, Shi Y, Zhang X, Xin X, Qi L, Guo H, Li J, Yang S. PIF3 is a negative regulator of the CBF pathway and freezing tolerance in Arabidopsis. Proc Natl Acad Sci USA. 2017;114(32):E6695–702.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  84. Wang S, Liang D, Li C, Hao Y, Ma F, Shu H. Influence of drought stress on the cellular ultrastructure and antioxidant system in leaves of drought-tolerant and drought-sensitive apple rootstocks. Plant Physiol Biochem. 2012;51:81–9.

    Article  PubMed  CAS  Google Scholar 

  85. Mahajan S, Tuteja N. Cold, salinity and drought stresses: an overview. Arch Biochem Biophys. 2005;444(2):139–58.

    Article  PubMed  CAS  Google Scholar 

  86. Ravindra KH. Molecular Cloning and expression of cor (Cold-Regulated) genes in Arabidopsis thaliana. Plant Physiol. 1990;93(3):1246–52.

    Article  Google Scholar 

  87. Thomashow MF. Genes induced during cold acclimation in higher plants. Advances in low-temperature biology. 1993; 2, 183–210.

  88. Chen Y, Jiang J, Chang Q, Gu C, Song A, Chen S, Dong B, Chen F. Cold acclimation induces freezing tolerance via antioxidative enzymes, proline metabolism and gene expression changes in two chrysanthemum species. Mol Biol Rep. 2014;41(2):815–22.

    Article  PubMed  CAS  Google Scholar 

  89. Niyogi KK. PHOTOPROTECTION REVISITED: Genetic and Molecular Approaches. Annual review of plant physiology and plant molecular biology. 1999; 50, 333–359.

  90. Apel K, Hirt H. Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu Rev Plant Biol. 2004;55:373–99.

    Article  PubMed  CAS  Google Scholar 

  91. Asada K. Production and scavenging of reactive oxygen species in chloroplasts and their functions. Plant Physiol. 2006;141:391–6.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  92. Tripathy BC, Oelmüller R. Reactive oxygen species generation and signaling in plants. Plant Signal Behav. 2012;7:1621–33.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  93. Fang Y, Liao K, Du H, Xu Y, Song H, Li X, Xiong L. A stress-responsive NAC transcription factor SNAC3 confers heat and drought tolerance through modulation of reactive oxygen species in rice. J Exp Bot. 2015;66(21):6803–17.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  94. Nuruzzaman M, Sharoni AM, Kikuchi S. Roles of NAC transcription factors in the regulation of biotic and abiotic stress responses in plants. Front Microbiol. 2013;4:248.

    Article  PubMed  PubMed Central  Google Scholar 

  95. Ahmad P, Jaleel CA, Salem MA, Nabi G, Sharma S. Roles of enzymatic and nonenzymatic antioxidants in plants during abiotic stress. Crit Rev Biotechnol. 2010;30(3):161–75.

    Article  PubMed  CAS  Google Scholar 

  96. Matsumura T, Tabayashi N, Kamagata Y, Souma C, Saruyama H. Wheat catalase expressed in transgenic rice can improve tolerance against low temperature stress. Physiol Plant. 2010;116(3):317–27.

    Article  Google Scholar 

  97. Badawi GH, Kawano N, Yamauchi Y, Shimada E, Sasaki R, Kubo A, Tanaka K. Over-expression of ascorbate peroxidase in Tobacco chloroplasts enhances the tolerance to salt stress and water deficit. Physiol Plant. 2004;121(2):231–8.

    Article  PubMed  CAS  Google Scholar 

  98. Eltayeb AE, Kawano N, Badawi GH, Kaminaka H, Sanekata T, Shibahara T, Inanaga S, Tanaka K. Overexpression of monodehydroascorbate reductase in transgenic Tobacco confers enhanced tolerance to ozone, salt and polyethylene glycol stresses. Planta. 2007;225(5):1255–64.

    Article  PubMed  CAS  Google Scholar 

  99. Kim YH, Kim CY, Song WK, Park DS, Kwon SY, Lee HS, Bang JW, Kwak SS. Overexpression of sweetpotato swpa4 peroxidase results in increased hydrogen peroxide production and enhances stress tolerance in Tobacco. Planta. 2008;227(4):867–81.

    Article  PubMed  CAS  Google Scholar 

  100. Dai F, Huang Y, Zhou M, Zhang G. The influence of cold acclimation on antioxidative enzymes and antioxidants in sensitive and tolerant barley cultivars. Biol Plant. 2009;3:257–62.

    Article  Google Scholar 

  101. Janda T, Szalai G, Rios-Gonzalez K, Veisz O, Pa´ldi E. Comparative study of frost tolerance and antioxidant activity in cereals. Plant Sci. 2003;64:301–6.

    Article  Google Scholar 

  102. Solte´sz A, Tı´ma´r I, Vashegyi I, To´th B, Kellos T, Szalai G, Va´gu´jfalvi A, Kocsy G, Galiba G. Redox changes during cold acclimation affect freezing tolerance but not the vegetative/ reproductive transition of the shoot apex in wheat. Plant Biol. 2011;13:757–66.

    Article  PubMed  Google Scholar 

  103. Gupta AS, Heinen JL, Holaday AS, Burke JJ, Allen RD. Increased resistance to oxidative stress in transgenic plants that overexpress chloroplastic Cu/Zn superoxide dismutase. Proc Natl Acad Sci USA. 1993;90(4):1629–33.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  104. Wang YJ, Wisniewski M, Meilan R, Cui MG, Webb R, Fuchigami L. Overexpression of cytosolic ascorbate peroxidase in tomato confers tolerance to chilling and salt stress. J Am SocHortic Sci. 2005;130:167–73.

    CAS  Google Scholar 

  105. Turan O, Ekmekci Y. Activities of photosystem II and antioxidant enzymes in chickpea (Cicer arietinum L.) cultivars exposed to chilling temperatures. Acta Physiol Plant. 2011;33:67–78.

    Article  CAS  Google Scholar 

  106. Zhang Y, Luo Y, Hou YX, Jiang H, Chen Q, Tang HR. Chilling acclimation induced changes in the distribution of H2O2 and antioxidant system of strawberry leaves. Agric J. 2008;3:286–91.

    Google Scholar 

  107. Ben Abdallah M, Methenni K, Nouairi I, Zarrouk M, Youssef NB. Drought priming improves subsequent more severe drought in a drought-sensitive cultivar of olive cv. Chétoui Scientia Horticulturae. 2017;221:43–52.

    Article  PubMed  Google Scholar 

  108. Chen T, Li W, Hu X, Guo J, Liu A, Zhang BA, Cotton MYB. Transcription factor, GbMYB5, is positively involved in Plant Adaptive Response to Drought stress. Plant Cell Physiol. 2015;56(5):917–29.

    Article  PubMed  CAS  Google Scholar 

  109. Vendruscolo EC, Schuster I, Pileggi M, Scapim CA, Molinari HB, Marur CJ, Vieira LG. Stress-induced synthesis of proline confers tolerance to water deficit in transgenic wheat. J Plant Physiol. 2007;164:1367–76.

    Article  PubMed  CAS  Google Scholar 

  110. Ruiz JM, Sanchez E, Garcia PC, Lopez-Lefebre LR, Rivero RM, Romero L. Proline metabolism and NAD kinase activity in greenbean plants subjected to cold-shock. Phytochemistry. 2002;59:473–8.

    Article  PubMed  CAS  Google Scholar 

  111. Gilmour SJ, Zarka DG, Stockinger EJ, Salazar MP, Houghton JM, Thomashow MF. Low temperature regulation of the Arabidopsis CBF family of AP2 transcriptional activators as an early step in cold-induced COR gene expression. The Plant Journal: For cell and Molecular Biology. 1998;16(4):433–42.

    Article  PubMed  CAS  Google Scholar 

  112. Liu Q, Kasuga M, Sakuma Y, Abe H, Miura S, Yamaguchi-Shinozaki K, Shinozaki K. Two transcription factors, DREB1 and DREB2, with an EREBP/AP2 DNA binding domain separate two cellular signal transduction pathways in drought- and low-temperature-responsive gene expression, respectively, in Arabidopsis. Plant Cell. 1998;10(8):1391–406.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  113. Shinwari ZK, Nakashima K, Miura S, Kasuga M, Seki M, Yamaguchi-Shinozaki K, Shinozaki K. An Arabidopsis gene family encoding DRE/CRT binding proteins involved in low-temperature-responsive gene expression. Biochem Biophys Res Commun. 1998;250(1):161–70.

    Article  PubMed  CAS  Google Scholar 

  114. Kidokoro S, Watanabe K, Ohori T, Moriwaki T, Maruyama K, Mizoi J, Myint Phyu Sin Htwe N, Fujita Y, Sekita S, Shinozaki K, Yamaguchi-Shinozaki K. Soybean DREB1/CBF-type transcription factors function in heat and drought as well as cold stress-responsive gene expression. The Plant Journal: For cell and Molecular Biology. 2015;81(3):505–18.

    Article  PubMed  CAS  Google Scholar 

  115. Steponkus PL, Uemura M, Joseph RA, Gilmour SJ, Thomashow MF. Mode of action of the COR15a gene on the freezing tolerance of Arabidopsis thaliana. Proc Natl Acad Sci USA. 1998;95(24):14570–5.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  116. Wang H, Datla R, Georges F, Loewen M, Cutler AJ. Promoters from kin1 and cor6.6, two homologous Arabidopsis thaliana genes: transcriptional regulation and gene expression induced by low temperature, ABA, osmoticum and dehydration. Plant Mol Biol. 1995;28(4):605–17.

    Article  PubMed  CAS  Google Scholar 

  117. Han Y, Xiang L, Song Z, Lu S. Overexpression of SgDREB2C from Stylosanthes guianensis leads to increased Drought Tolerance in Transgenic Arabidopsis. Int J Mol Sci. 2022;24(7):3520.

    Article  Google Scholar 

Download references

Acknowledgements

We are grateful to Li Baocheng, the Vice President of Xinjiang Academy of Agricultural Reclamation, for providing Xinjiang Land Cotton XinLuzao 17 seeds. We also thank the reviewers and editors for their careful reading and helpful comments on this paper.

Funding

This work was supported by the High-level Talents Research Start-up Fund of Science and Technology Research and Development Program of Shihezi University (RCZK202045), the Establishment of an efficient genetic transformation system for self-breeding land cotton varieties in Xinjiang and creation and demonstration of high insect-resistant cotton germplasm (KY2023GG03), the Establishment of an efficient genetic transformation system for land cotton and creation and demonstration of new germplasm (2023C13) and the Independent Research project of Shihezi University (ZZZC202082B).

Author information

Authors and Affiliations

Authors

Contributions

Mei Wang and Lepeng Wang were the experimental designers and study executors of the study, completed the data analysis, and wrote the first draft of the manuscript. Xiangxue Yu and Jingyi Zhao carried out the experimental results and data analysis. Zhijia Tian, Xiaohong Liu, and Guoping Wang participated in the experimental data inspection and data investigation. Xinyong Guo and Li Zhang were the project planners and principals, guiding the paper’s writing and revision. All authors read and agreed to the final text.

Corresponding author

Correspondence to Xinyong Guo.

Ethics declarations

Ethics approval and consent to participate

The plant material Saussurea involucrata (Matsum. & Koidz) used in this study was obtained from the Key Laboratory of Agricultural Biotechnology, College of Life Sciences, Shihezi University, Xinjiang, China. All the Saussurea involucrata (Matsum. & Koidz) used in this experiment were laboratory histoculture seedlings, not collected from endangered materials or species. All the procedures were carried out following the relevant guidelines.

Consent for publication

Not applicable.

Competing interests

The authors declare no competing interests.

Additional information

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary Material 1

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, M., Wang, L., Yu, X. et al. Enhancing cold and drought tolerance in cotton: a protective role of SikCOR413PM1. BMC Plant Biol 23, 577 (2023). https://doi.org/10.1186/s12870-023-04572-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12870-023-04572-6

Keywords