Next Article in Journal
WGS Data Collections: How Do Genomic Databases Transform Medicine?
Previous Article in Journal
New Dinuclear Macrocyclic Copper(II) Complexes as Potentially Fluorescent and Magnetic Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Functional Characterization of Lobularia maritima LmTrxh2 Gene Involved in Cold Tolerance in Tobacco through Alleviation of ROS Damage to the Plasma Membrane

1
Centre of Biotechnology of Sfax, Biotechnology and Plant Improvement Laboratory, University of Sfax, B.P “1177”, Sfax 3018, Tunisia
2
Plant Production Department, College of Food and Agricultural Sciences, King Saud University, P.O. Box 2460, Riyadh 11451, Saudi Arabia
3
Biology Department, College of Sciences and Arts Muhayil Assir, King Khalid University, Abha 61421, Saudi Arabia
4
Department of Chemistry, Faculty of Sciences and Arts in Balgarn, University of Bisha, Bisha 61922, Saudi Arabia
5
Department of Environmental Sciences and Nutrition, Higher Institute of Applied Sciences and Technology of Mahdia, University of Monastir, Mahdia 5100, Tunisia
6
Faculty of Horticulture, Institute of Horticulture, Slovak University of Agriculture, Tr. A. Hlinku 2, 949 76 Nitra, Slovakia
7
Department of Bioenergy, Food Technology and Microbiology, Institute of Food Technology and Nutrition, University of Rzeszow, 4 Zelwerowicza St, 35601 Rzeszow, Poland
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(3), 3030; https://doi.org/10.3390/ijms24033030
Submission received: 10 December 2022 / Revised: 29 January 2023 / Accepted: 1 February 2023 / Published: 3 February 2023
(This article belongs to the Section Molecular Plant Sciences)

Abstract

:
Cold stress is a key environmental factor affecting plant growth and development, crop productivity, and geographic distribution. Thioredoxins (Trxs) are small proteins that are ubiquitously expressed in all organisms and implicated in several cellular processes, including redox reactions. However, their role in the regulation of cold stress in the halophyte plant Lobularia maritima remains unknown. We recently showed that overexpression of LmTrxh2, which is the gene that encodes the h-type Trx protein previously isolated from L. maritima, led to an enhanced tolerance to salt and osmotic stress in transgenic tobacco. This study functionally characterized the LmTrxh2 gene via its overexpression in tobacco and explored its cold tolerance mechanisms. Results of the RT-qPCR and western blot analyses indicated differential temporal and spatial regulation of LmTrxh2 in L. maritima under cold stress at 4 °C. LmTrxh2 overexpression enhanced the cold tolerance of transgenic tobacco, as evidenced by increased germination rate, fresh weight and catalase (CAT), superoxide dismutase (SOD) and peroxidase (POD) activities; reduced malondialdehyde levels, membrane leakage, superoxide anion (O2), and hydrogen peroxide (H2O2) levels; and higher retention of chlorophyll than in non-transgenic plants (NT). Furthermore, the transcript levels of reactive oxygen species (ROS)-related genes (NtSOD and NtCAT1), stress-responsive late embryogenis abundant protein 5 (NtLEA5), early response to dehydration 10C (NtERD10C), DRE-binding proteins 1A (NtDREB1A), and cold-responsive (COR) genes (NtCOR15A, NtCOR47, and NtKIN1) were upregulated in transgenic lines compared with those in NT plants under cold stress, indicating that LmTrxh2 conferred cold stress tolerance by enhancing the ROS scavenging ability of plants, thus enabling them to maintain membrane integrity. These results suggest that LmTrxh2 promotes cold tolerance in tobacco and provide new insight into the improvement of cold-stress resistance to cold stress in non-halophyte plants and crops.

1. Introduction

Plants are sessile in nature and thus face various environmental stresses throughout their life cycles, including abiotic (such as drought, heat, salinity, and cold) and biotic (such as fungi, bacteria, and viruses) factors. Among these stresses, cold causes losses in crop yields, making it a limiting factor in agricultural production [1,2]. To adapt and tolerate cold stress, plants have developed sophisticated defense mechanisms by altering biochemical and physiological processes, including the accumulation of cryoprotective polypeptides (e.g., COR15a) and osmolytes (e.g., soluble sugars and proline) [1,3,4]. Under cold stress conditions, plants generate excessive reactive oxygen species (ROS), including the superoxide radical (O2.−), hydrogen peroxide (H2O2), and singlet oxygen (1O2) [5,6], which accelerate the peroxidation of membrane lipids and even lead to cell death when they are not timely scavenged [7]. Therefore, ROS homeostasis is critical for the normal growth and development of plants. However, the most stable form of ROS, hydrogen peroxide (H2O2), diffuses freely in plants and plays a crucial role as a signaling molecule in various physiological processes [8]. Nevertheless, to regulate ROS concentrations, plant cells are also equipped with several antioxidant enzymes and proteins in various subcellular organelles, such as ascorbate peroxidase, catalase, glutathione peroxidase, peroxiredoxin (Prx), glutaredoxin, and thioredoxin (Trx) [9,10]. Numerous studies have indicated that TFs have significant roles in regulating various biological processes to protect plant cells against cold induced damage [4,11]. NtbHLH123 expression was induced by cold stress, suggesting that NtbHLH123 may have a regulatory role during cold stress [12]. Several studies have shown that the overexpression of ICE1, which results in insensitivity to chilling and freezing stresses, positively regulates CBF gene expression under cold stress [13]. In addition, CBF genes are induced by cold, and their overexpression enhances cold tolerance in transgenic Arabidopsis, rice, tobacco, tomato, and apple [14,15,16].
Thioredoxins represent a separate group of low molecular weight redox proteins that are controlled by a family multigene in plants [17]. Trxs have several crucial roles in many cellular processes in plants, such as the germination of seeds, the assimilation of carbon, the metabolism of lipids, phytohormone and redox signaling, and the stress response [18,19,20,21,22,23,24]. In plants, Trxs are implicated in the control of transcription factors, activate ribonucleotide reductase, enhance photosynthetic efficiency, and regulate enzyme activities [25]. Trxs are implicated in oxidative stress responses in plants by (1) repairing oxidative proteins (such as iron–sulfur protein and DNA damage repair related proteins) [21]; (2) activating the activity of protecting enzymes in the antioxidant system [26]; and (3) acting as regulators of scavenging mechanisms or signaling pathways in the antioxidant network [23,27]. In addition, the involvement of these same proteins in cold stress tolerance in rice [28] and potatoes [29] was also reported. Park et al. [30] proved that the cytosolic Trx-h2 protein from Arabidopsis relays the external cold stress signal to downstream cold defense signaling cascades through its protein disulfide reductase function. Lee et al. [31] showed that Arabidopsis disulfide reductase, Trx-h2, plays the role of a RNA chaperone under cold stress, which boosts the plant’s tolerance. Additionally, Lee et al. [32] proved that Arabidopsis Trx-hs regulates target proteins’ redox status via exchanging disulfide bonds between thioredoxin and the targets, especially Trx-h2, which regulates the redox status and induces the structural switching of C-repeat binding factors (CBFs). Recently, Kopeć et al. [33] reported that light plays a significant role in acquiring cold/freezing tolerance, apart from low temperatures. A low/freezing temperature disrupts the balance between light-harvesting and the subsequent processing of this energy through dark reaction biochemistry, which is downregulated under cold. Thus, the unused excess excitation energy can lead to the formation of reactive oxygen species which may interact with thioredoxins, among others. Wu et al. [34] reported that MaTrx12 from Musa acuminata plays an important role in the chilling tolerance of harvested banana fruit, possibly by regulating redox homeostasis. Trx-h1 and Trx-h4 from Arabidopsis thaliana are involved in cell cycle control and proliferation [35]. Trx-h3 is known to play a crucial role in the plant’s defense response against oxidative stress [36,37]. Trx-h5 is important for the plant’s immune response and the structural and functional regulation of the nonexpressor of pathogenesis-related genes 1 (NPR1) [38,39]. Various studies have shown that plant Trx-hs genes respond to various external stresses [40,41,42,43,44]; however, the responses are species-dependent. The specific function of several Trx-h isoforms remains unclear, especially the implication of Trx-h2 in response to cold stress in the halophyte plant L. maritima. Based on the hypothesis that halophytes could be a valuable source of stress-adaptive genes, we previously reported that the AlSAP gene and its promoter isolated from the halpoyte plant Aeluropus littoralis were induced by high- and low-temperature stress [45,46]. The constitutive expression of AlSAP in tobacco and rice has resulted in an enhanced tolerance to cold, heat, drought, salinity, and oxidative stresses [45,47]. Recently, AlTMP1 gene encoding the plasma membrane protein 1 from A. littoralis was found to be inducible by low-temperature (4 °C), and its overexpression in transgenic tobacco conferred tolerance to freezing (−20 °C for 2 h) and heat (55 °C for 2 h) stresses [48]. Previously, we isolated and characterized the first Trxh2 gene of the halotolerant plant L. maritima, which was designated as LmTrxh2 and deposited in the GenBank databases (accession number: MW589650) [49]. LmTrxh2 is a thioredoxin H-type protein characterized by the presence of a conserved TRX domain and catalytic Trx motif CGPC, which exhibits 91.1% similarity with Arabidopsis thaliana [49]. The LmTrxh2 gene is inducible by salt and osmotic and oxidative stresses in L. maritima localized in the plasma membrane, and its overexpression in transgenic tobacco plants resulted in an enhanced tolerance to salt and osmotic stresses [49]. The aim of this work was therefore to study the role of LmTrxh2 under cold stress. We functionally characterized a differentially expressed LmTrxh2 gene in L. maritima via inducible expression pattern analysis. LmTrxh2-overexpressed tobacco plants were investigated in cold stress conditions, and the experimental results showed that an overexpression of LmTrxh2 in tobacco plants conferred them tolerance to cold stress. Furthermore, these results might help in clarifying the response mechanisms of LmTrxh2 under cold stress condition.

2. Results

2.1. LmTrhx2 Transcription Is Highly Induced by Cold Stress

To identify the cold stress response of LmTrhx2, we exposed the L. maritima seedlings to cold stress and used RT-qPCR to analyze the relative expression level of LmTrhx2. When stressed at 4 °C, we observed that the LmTrhx2 mRNA level significantly increased in the leaves, with a 2.5-fold expression at 1 h, while it reached a peak (6-fold) at 3 h, and it decreased slightly at 24 h (Figure 1A). LmTrhx2 was also remarkably induced in the root within 1 h of the treatment and reached a peak at 12 h (4-fold induction) (Figure 1A). A western blot analysis technique was conducted to determine whether the transcription level of LmTrxh2 was correlated with the accumulation of protein levels in leaves and roots tissues that were treated or not treated with cold stress. Indeed, the immunoblots analysis confirmed the above obtained results of LmTrxh2 expression and suggested that the LmTrxh2 protein level was abundant in the leaves and roots of L. maritima under cold stress (Figure 1B). Hence, these results suggest that cold stress induces LmTrxh2 transcription and protein accumulation in L. maritima.

2.2. Transgenic Tobacco Plants Harboring LmTrxh2 Gene Enhanced Cold Tolerance

We have used three previously established homozygous LmTrxh2 tobacco lines (Tr1, Tr2, and Tr3) that constitutively express LmTrxh2 gene, as reported previously [49]. Then, their tolerance responses to cold stress (4 °C) were investigated. Seeds of Tr1, Tr2, Tr3 and NT lines were sown in vitro on MS media under 25 °C or 4 °C temperature, and the germination rates were determined afterwards. In fact, transgenic and NT seeds had the same germination levels under 25 °C, and all seeds had germinated after 10 days.
As illustrated in Figure 2A, the three transgenic lines were phenotypically characterized by more vigorous growth and greener leaves and significantly higher germination rates compared with NT plants under cold stress (4 °C). Indeed, 58–64% of seeds from the previously reported three transgenic lines had germinated compared to only 30% of the NT seeds. Therefore, this indicates that Lmtrxh2 overexpression in tobacco plants may have enhanced the seeds’ germination tolerance to cold stress (Figure 2B). As shown in Figure 2C, transgenic lines showed better growth parameters compared with the NT plants. In fact, fresh weight accumulation, shoot length, and root length were significantly higher in the transgenic lines (Tr1, Tr2, and Tr3) than those in the NT plants (Figure 2D and Figure S1). We further measured the total chlorophyll content in the three transgenic lines and found that following cold treatment, the total chlorophyll levels of Tr1 (12.97 µg/mL), Tr2 (11.50 µg/mL), and Tr3 (12.38 µg/mL) transgenic plants were higher than the NT plants (4.96 µg/mL) (Figure 2E).

2.3. LmTrxh2 Overexpression Alleviated Free Radical Damage on the Membrane under Cold Stress

Seedlings that were 45 days old were used to further investigate cold stress tolerance in tobacco plants. No morphological differences were observed between transgenic and NT plants before cold treatment (Figure 3A); however, when the seedlings were subjected to cold stress (4 °C) for 7 days, serious cold injuries were recorded in the NT plants compared with the transgenic plants. Following a 15-day recovery period under 25 °C, most of the NT plants died, while the majority of the LmTrxh2-overexpressing plants survived (Figure 3A).
One serious drawback of cold stress exposure is the generation of ROS, which stimulates oxidative stress and consequently results in a damaging effect on the different cell components. Furthermore, malonaldehyde (MDA) is an important indicator regarding the degree of lipid peroxidation which is commonly used to estimate membrane damage under oxidative stress [50]. Thus, we evaluated the accumulation of H2O2 and the content of lipid peroxidation in LmTrxh2 transgenic seedlings that were grown under normal or cold stress conditions. Although there was no difference between the MDA (Figure 3B) and H2O2 (Figure 3C) contents in NT and transgenic plants under normal conditions, the transgenic plants had significantly lower MDA and H2O2 contents than the NT plants under cold stress. The rate of O2 accumulation in the leaves of transgenic lines and NT plants was also quantified spectrophotometrically (Figure 3D). It increased after 48 h of cold stress in three transgenic lines but at significantly lower levels than the NT plants. We can conclude from this data that the overexpression of LmTrxh2 had an inhibitory effect on the accumulation of ROS induced by cold stress in plants. We further measured the El to determine the membrane integrity of the plants under cold stress. No significant differences were found between transgenic and NT leaves under normal conditions, while El values were significantly lower in transgenic leaves under cold stress (Figure 3E).
Since the tobacco plants overexpressing LmTrxh2 had enhanced cold tolerance, we subjected transgenic and NT leaves to NBT and DAB staining in order to detect O2 and H2O2 production in young seedlings treated with cold stress for 48 h.
As shown in Figure 4A,B, under normal conditions, no significant differences were found in the accumulation of O2 and H2O2 between the leaves of LmTrxh2 transgenic tobacco and those of the NT plants. However, under cold treatment, the NBT and DAB staining were stronger and intensified in all lines. Nevertheless, in the LmTrxh2-overexpressing lines, exposure to cold stress led to less accumulation of O2 and H2O2 compared with the NT plants.
Our results thus suggest that the accumulation of ROS was higher in the NT plants compared with the transgenic plants, and this could be due to the activation of three key antioxidant enzymes (CAT, POD, and SOD), which are crucial in ROS scavenging and ROS cellular homeostasis [51]. We hence quantified their levels before and after cold stress treatments and found that the levels of the expression of the three proteins were similar between the NT and transgenic plants under normal conditions. Following cold treatment, however, LmTrxh2-overexpressing plants showed an enhancement in the activities of all antioxidant enzymes. In contrast, the levels of these same proteins in the NT plants only slightly increased (Figure 4C–E). These data suggest that LmTrxh2 overexpression could alleviate membrane damage by enhancing the free radical scavenging ability of tobacco, thus promoting its cold stress tolerance.

2.4. LmTrxh2 Altered the Expression Levels of the Cold-Responsive and ROS Scavenging-Related Genes

The enhanced cold tolerance in transgenic tobacco plants was conferred by the LmTrxh2 overexpression-induced enhancement of the free radical scavenging ability. Consequently, an induction in the expression of ROS-related genes led to an enhanced production of enzymes that were implicated in direct ROS detoxification in the NT and transgenic plants before and after cold treatment. The transcript levels of NtSOD and NtCAT1 were similar between NT and transgenic lines under normal conditions but were significantly upregulated in the transgenic lines compared with the NT under cold treatment. This suggests that LmTrxh2 could be a key regulator gene that may be found upstream of some ROS-related genes. Therefore, the overexpression of LmTrxh2 could result in the activation of the expression of several ROS-related genes that may help plants positively cope with environmental stresses. Additionally, we investigated several transcript levels of many cold stress defensive proteins (NtDREB1A, NtCOR15A, NtCOR47, NtKIN1, NtLEA5, and NtERD10D). The results showed that following cold stress, the expression levels of the cold-related genes enhanced 4.6- to 5.6-fold in LmTrxh2 transgenic lines compared with NT plants (Figure 5). These results, therefore, suggest that the overexpression of LmTrxh2 may increase the expression levels of the mRNA of ROS scavenging-related and stress-responsive genes following cold stress.

3. Discussion

In this study, the LmTrxh2 gene from L. maritima was functionally demonstrated to be implicated in the regulation of the plant’s cold tolerance. This effect was attributed to an enhanced H2O2-scavenging capacity and intact membrane integrity maintenance induced by LmTrxh2 overexpression and the upregulation of cold stress defensive genes. In a previous report, we showed that the LmTrxh2 gene enhanced salt and osmotic stress tolerance to tobacco through the regulation of redox homeostasis [49]. In this study, our results showed that the expression of the LmTrxh2 gene was differentially induced in the leaves and roots of L. maritima by a low temperature, suggesting that LmTrxh2 may have a regulatory role during cold stress. Duan et al. [52] showed that the expression level of the MsTRX gene was significantly altered by different abiotic stresses, such as cold, drought, and salt in alfalfa. Furthermore, mRNA level of Arabidopsis NADPH-dependent thioredoxin reductase C (AtNTRC) was also induced in response to cold stress, which confers freezing and cold shock tolerance to plants [53]. In addition, several Trxs have been found to accumulate specifically in certain tissues. Similar to our results, the expressions of LjTrxf, LjTrxm1, LjTrxm4, and LjTrxx in Lotus japonicas were higher in leaves than in roots [54]. For instance, pea thioredoxin PsTRXf1 was found in early seedlings, leaves, roots, stems, and flowers, while the expression of PsTRXm1 was restricted to leaves, roots, seeds, and flowers [55]. Trxs h2 (MtTrx1) and h6 (MtTrx31) were abundant in cotyledons of the legume Medicago truncatula [56]. Similarly, Zhang et al. [23] reported that an h-type Trx gene in Oryza sativa (OsTrxh1) was induced by salt stress in root tissues and can influence the content of ROS by regulating the apoplastic ascorbate system or other antioxidant enzymatic activities. An upregulation of a Glycina max h-type Trx gene (GmTrxh) in the roots and infected cells of mature nodules after inoculation with a nodulating bacterium was observed, and this gene was found to contain an antioxidant responsive element (ARE)-like site in its 5′ upstream region, and its cDNA conferred tolerance to H2O2 in a yeast thioredoxin mutant [57]. Collectively, we conclude from these findings that the modulation of the LmTrxh2 gene in leaves and roots might be associated with cold tolerance mechanisms in L. maritima.
TRX genes were reported to be crucial for the regulation of biological mechanisms implicated in the protection of plants against cold-induced damages [40,43]. In the present work, we showed that following the stress cold tolerance assay, the transgenic tobacco plants were more capable of tolerating cold stress compared with the NT ones. This was proven through the measurement of the survival rate, EL (%), MDA, H2O2, and total chlorophyll levels, which were in accordance with the phenotypic observation that suggested that the overexpression of LmTrxh2 enhanced cold tolerance. Likewise, Park et al. [37] showed that the overexpression of the Arabidopsis Trx-h3 gene conferred heat shock tolerance in plants. Additionally, in harvested banana, the gene MaTrx12 was shown to be involved in cold tolerance [34], and the overexpression of AtTrx-h2 gene was found to confer resistance to salt stress in Brassica napus [58]. Similarly, the overexpression of the gene GhTRX134 in Arabidopsis was reported to enhance drought and salt tolerance as well as resistance to oxidative stress [59]. Following exposure to cold stress, the balance between ROS production and clearance is known to be disrupted [5,51], which results in multiple cytological effects that may include plasma membrane lipid peroxidation [60]. In fact, plasma membrane integrity is directly correlated with the degree of membrane lipid peroxidation. Therefore, this implies that efficient scavenging of ROS in the plasma membrane may endow plants with higher cold tolerance. Our subcellular location analysis indicated that LmTrxh2 is localized at the membrane and cytoplasm [49], suggesting its role in the degradation of over-accumulated H2O2 in the membrane under cold stress. Nevertheless, the regulation of the ROS system under different abiotic stresses remains dependent on the ROS-scavenging systems, especially through the antioxidant enzymes SOD, CAT, and POD [5,61]. These enzymes are regulated by TRX genes [62] and were found to be highly expressed under cold stress in the transgenic plants compared with the NT ones. Their scavenging activity was also indirectly proven by the reduction of H2O2 and MDA contents under cold stress in the transgenic plants.
In the present study, we further showed that the two ROS-scavenging related genes NtSOD and NtCAT1 were highly expressed in the LmTrxh2-overexpressing plants compared with the NT plants under cold stress. Indeed, the level of expression of these two genes correlated with the enhanced activities of the three previously mentioned antioxidant enzymes. Therefore, we suggest that the overexpression of LmTrxh2 could enhance cold tolerance due to a better ROS-scavenging system.
One of the regulator genes that plays a crucial role in the cold adaptation of plants is the cold-regulated (COR) gene. Studies have shown that about 10–20% of the COR gene is regulated by C-repeat binding factors (CBFs) [63,64,65]; for example, expressions of 24 COR genes are regulated by CBF-dependent cold-induced zinc finger transcription factor ZAT12 [66]. Cold stress also induces the expression of late embryogenesis abundant (LEA), COR15A, and COR15B proteins, which can interact with membranes and act as membrane protectants [67]. The expression of the COR15A gene that encodes a hydrophilic protein has been reported to improve the cold tolerance of the chloroplasts of domestic plants [68]. In addition, COR47 is also closely related to cold tolerance because it acts as an anti-dehydrating agent that prevents excessive dehydration of plant cells due to low temperatures [69]. Interestingly, RT-qPCR analysis before and after cold stress showed that the mRNA levels of cold-responsive genes, such as DREB1A (dehydration response element B1A), COR15A (cold-regulated 15A), COR47 (cold-regulated 47), KIN1 (stress-responsive protein KIN1), NtERD10D (early response to dehydration 10D), and NtLEA5 (late embryogenesis abundant protein) or their homologs respond to abiotic stresses [70,71]. They were found to be strongly expressed in the transgenic plants compared with the NT ones, suggesting minimal damage to those plants, which is evidenced by high membrane integrity under cold stress. In future studies, we plan to unravel whether the LmTrxh2 gene was directly implicated in the regulation of the stress-responsive genes to improve plant cold-stress resistance.
This study functionally characterized the LmTrxh2 gene from L. maritima, which specifically enhanced the tolerance of tobacco plants to cold stress. This specific response could be attributed to its role in ROS scavenging and the upregulation of stress-responsive genes to maintain plasma membrane integrity.

4. Materials and Methods

4.1. Plant Materials and Growth Conditions

Sample of seeds of L. maritima were taken from saline marshes near the region of Chebba, Mahdia in Tunisia. They were then dried and stored at 4 °C until further use. Sterilization and germination procedures of the seeds were done following protocols previously described by Ben Saad et al. [72]. Seedlings were then left to grow for four weeks in a nutrient solution, as reported elsewhere [73], and they were also subjected to cold treatment. The treatment consisted of exposing the plants to a low temperature (4 °C) under an illumination scheme of 16 h/8 h (light/dark) with a light intensity of 35 μmol m−2 s−1. Thereafter, plants tissues were collected at 1, 3, 6, 12, 24, and 48 h and were immediately frozen in liquid nitrogen and preserved at −80 °C for further analyses.

4.2. RNA Isolation, cDNA Synthesis, and Quantitative Reverse Transcription PCR Analysis

In order to analyze the expression of LmTrxh2 transcript in L. maritima, the leaves and roots of L. maritima were collected at 1, 3, 6, 12, 24, and 48 h of cold stress (4 °C). RNA isolation and cDNA synthesis were performed according to our previously published method [49]. Briefly, RNA from L. maritima and transgenic tobacco plants were extracted using the standard TRIzol method. The obtained RNA was treated with DNase I (MBI Fermentas, Hanover, MD, USA) at 37 °C for 15 min to remove any remaining genomic DNA and used as a template for cDNA synthesis [74]. RT-qPCR were conducted as described by Ben Saad et al. [74]. The forward and reverse primers (qLmTrxh2-F and qLmTrxh2-R, respectively) were used for the RT-qPCR analysis of LmTrxh2 expression. The UBQ10 gene (UBQ10-F and UBQ10-R) was used as an internal control, and the relative expression of the target was calculated using the 2−ΔΔCT method [75]. Three technical replicates and three biological replicates were used for each treatment. The primer sequences used for the RT-qPCR are listed in Supplementary Table S1. The RT-qPCR was also used to determine the transcript accumulation of LmTrxh2 in homozygous transgenic tobacco lines (Tr1, Tr2, and Tr3) and NT tobacco plants and to monitor the expression levels of ROS-related (NtSOD and NtCAT1) and cold-response genes (NtDREB1A, NtCOR15A, NtCOR47, NtKIN1, NtLEA5 and NtERD10C) (Table S1) in Tr1, Tr2, and Tr3 transgenic lines exposed to 4 °C for 48 h. Total RNA isolation from tobacco leaf tissues, cDNA synthesis, and RT-qPCR using gene specific primers were performed as described above. The Actin gene (ACT-F and ACT-R) was used as an internal control (Table S1).

4.3. Western Blot Analysis

The Western blotting method was used to quantify the accumulation of LmTrxh2 in the tissues of the leaves and roots of L. maritima that were collected after 1, 3, 6, 12, 24, and 48 h of cold treatment (4 °C). Total proteins extraction, separation, and LmTrxh2 protein detection were conducted as previously described by Ben Saad et al. [76]. Monoclonal antibody to β-actin was purchased from Sigma Aldrich Co., St. Louis, MO, USA (Cat#A5441).

4.4. Cold Tolerance Analysis of the Transgenic Plants

The seeds of transgenic homozygous T3 generation tobacco lines (Tr1, Tr2, and Tr3) and NT plants were surface-disinfected and germinated on MS medium and incubated for 7 days at 25 ± 1 °C or 4 ± 1 °C. Their germination rates were recorded, and the images were captured two weeks after growth in a culture chamber under a 16 h/8 h light/dark cycle at 25 °C. After a two-week growth period, the fresh weight of the plants and the chlorophyll content of the leaves were determined. The total chlorophyll content in each sample was calculated after extraction in aqueous 80% acetone [77] using the following formulae which express [Chl a], [Chl b], and [Chls a + b] in µg/mL: [Chl a] = 12.70 × A663 − 2.69 × A645, [Chl b] = 22.90 × A645 − 4.68 × A663, and [Chls a + b] = 20.21 × A645 + 8.02 × A663. The A663 and A645 represent absorbance values read at 663 nm and 645 nm wavelengths, respectively. The assays were conducted in triplicate on independent seed lots.
Regarding the analysis of the grown seedlings, the plants were first put into plastic containers that were filled with both soil and sand (1:1) under controlled conditions (25 °C temperature, 16 h light/8 h dark, 70% humidity, and 200 µM m−2 s−1 light intensity). Then, the transgenic and NT transgenic tobacco plants were aged of 45 days. The 45-day-old tobacco plants of the NT and transgenic plants (Tr1, Tr2, and Tr3) were transferred to the growth chamber under the previously mentioned controlled conditions for 7 days. Thereafter, the cold-stressed plants were returned to the normal growth chamber to grow for 15 more days. Then, tissues of the treated plants (4 °C for 48 h) and the control plants were collected for biochemical (oxidative stress markers and antioxidant enzyme activities) analysis and electrolyte leakage (El). All the results were based on the average of three independent biological replicates.

4.5. Histochemical Staining Analysis

The hydrogen peroxide (H2O2) and superoxide radicals (O2) accumulation in the leaves were stained using DAB and NBT solution, as described by Ben Hsouna et al. [78]. Briefly, leaves that were stained with DAB or NBT were treated with 70% ethanol to remove the chlorophyll content and make the staining more visible. Then, the stained slides were observed and photographed with a Leica MZ FLIII binocular microscope (Leica Microsystems, Heerbrugg, Switzerland).

4.6. Determination of the Physiological Indexes

The activities of antioxidant enzymes superoxide dismutase (SOD), peroxidase (POD), catalase (CAT), and the content of malondialdehyde (MDA) and H2O2 were measured as previously described [79]. Superoxide content (O2) was quantitated as described by Chen et al. [80]. Furthermore, the El was measured following a previously published method [48]. Furthermore, the total chlorophyll content of the leaves was determined as described previously by Ben Romdhane et al. [72].

4.7. Statistical Analysis

Data were analyzed using the Statistical Analysis System software (IBM SPSS Statistics 21.0, IBM Corp., Armonk, NY, USA), and significant differences among treatments were determined using Duncan’s multiple range test (p < 0.05). The data were presented as means ± SD values of three biological replicates (n = 3).

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24033030/s1.

Author Contributions

Conceptualization, R.B.S.; methodology, R.B.S. and M.T.B.; formal analysis, R.B.S., A.B.H., W.M. and M.K.; resources, R.B.S. and M.K.; data curation, M.T.B., Y.C. and R.B.S.; writing—original draft preparation, R.B.S.; writing—review and editing, R.B.S., N.B. and W.B.R.; supervision, R.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

N.B. was funded through the Small Research Group Project from the Deanship of Scientific Research at King Khalid University under research grant number (R.G.P.1/295/43). This research was also partially funded by the Tunisian Ministry of Higher Education and Scientific Research (Program contract 2019–2022). The APC was funded by KEGA project no. 010SPU-4/2021.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data is contained within the article or supplementary material.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the Small Research Group Project under grant number (R.G.P.1/295/43).

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Chinnusamy, V.; Zhu, J.; Zhu, J.K. Cold stress regulation of gene expression in plants. Trends Plant Sci. 2007, 12, 444–451. [Google Scholar] [PubMed]
  2. Sanghera, G.S.; Wani, S.H.; Hussain, W.; Singh, N.B. Engineering Cold Stress Tolerance in Crop Plants. Curr. Genom. 2011, 12, 30–43. [Google Scholar] [CrossRef] [PubMed]
  3. Thomashow, M.F. PLANT COLD ACCLIMATION: Freezing Tolerance Genes and Regulatory Mechanisms. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1999, 50, 571–599. [Google Scholar] [CrossRef]
  4. Zhu, J.-K. Abiotic Stress Signaling and Responses in Plants. Cell 2016, 167, 313–324. [Google Scholar] [CrossRef]
  5. Choudhury, F.K.; Rivero, R.M.; Blumwald, E.; Mittler, R. Reactive oxygen species, abiotic stress and stress combination. Plant J. 2017, 90, 856–867. [Google Scholar] [CrossRef]
  6. Gilroy, S.; Białasek, M.; Suzuki, N.; Górecka, M.; Devireddy, A.R.; Karpiński, S.; Mittler, R. ROS, Calcium, and Electric Signals: Key Mediators of Rapid Systemic Signaling in Plants. Plant Physiol. 2016, 171, 1606–1615. [Google Scholar] [CrossRef]
  7. Gechev, T.S.; Van Breusegem, F.; Stone, J.M.; Denev, I.; Laloi, C. Reactive oxygen species as signals that modulate plant stress responses and programmed cell death. Bioessays 2006, 28, 1091–1101. [Google Scholar] [CrossRef]
  8. Veal, E.; Day, A. Hydrogen Peroxide as a Signaling Molecule. Antioxid. Redox Signal. 2011, 15, 147–151. [Google Scholar] [CrossRef]
  9. Hanschmann, E.M.; Godoy, J.R.; Berndt, C.; Hudemann, C.; Lillig, C.H. Thioredoxins, glutaredoxins, and peroxiredox-ins—Molecular mechanisms and health significance: From cofactors to antioxidants to redox signaling. Antioxid. Redox Signal. 2013, 19, 1539–1605. [Google Scholar] [CrossRef]
  10. Lee, E.S.; Kang, C.H.; Park, J.H.; Lee, S.Y. Physiological significance of plant peroxiredoxins and the structure-related and multifunctional biochemistry of peroxiredoxin 1. Antioxid. Redox Signal. 2018, 28, 625–639. [Google Scholar] [CrossRef]
  11. Shi, Y.; Ding, Y.; Yang, S. Cold Signal Transduction and its Interplay with Phytohormones during Cold Acclimation. Plant Cell Physiol. 2014, 56, 7–15. [Google Scholar] [CrossRef]
  12. Zhao, Q.; Xiang, X.; Liu, D.; Yang, A.; Wang, Y. Tobacco Transcription Factor NtbHLH123 Confers Tolerance to Cold Stress by Regulating the NtCBF Pathway and Reactive Oxygen Species Homeostasis. Front. Plant Sci. 2018, 9, 381. [Google Scholar] [CrossRef]
  13. Budhagatapalli, N.; Narasimhan, R.; Rajaraman, J.; Viswanathan, C.; Nataraja, K.N. Ectopic expression of AtICE1 and OsICE1 transcription factor delays stress-induced senescence and improves tolerance to abiotic stresses in tobacco. J. Plant Biochem. Biotechnol. 2015, 25, 285–293. [Google Scholar] [CrossRef]
  14. Hsieh, T.H.; Lee, J.T.; Charng, Y.Y.; Chan, M.T. Tomato plants ectopically expressing Arabidopsis CBF1 show enhanced re-sistance to water deficit stress. Plant Physiol. 2002, 130, 618–626. [Google Scholar]
  15. Ito, Y.; Katsura, K.; Maruyama, K.; Taji, T.; Kobayashi, M.; Seki, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Functional analysis of rice DREB1/CBF-type transcription factors involved in cold-responsive gene expression in transgenic rice. Plant Cell Physiol. 2006, 47, 141–153. [Google Scholar]
  16. Yang, W.; Liu, X.-D.; Chi, X.-J.; Wu, C.; Li, Y.-Z.; Song, L.-L.; Liu, X.-M.; Wang, Y.-F.; Wang, F.-W.; Zhang, C.; et al. Dwarf apple MbDREB1 enhances plant tolerance to low temperature, drought, and salt stress via both ABA-dependent and ABA-independent pathways. Planta 2010, 233, 219–229. [Google Scholar] [CrossRef]
  17. Meyer, Y.; Belin, C.; Delorme-Hinoux, V.; Reichheld, J.P.; Riondet, C. Thioredoxin and glutaredoxin systems in plants: Mo-lecular mechanisms, crosstalks, and functional significance. Antioxid. Redox Signal. 2012, 17, 1124–1160. [Google Scholar] [CrossRef]
  18. Collet, J.-F.; Messens, J.; Tomasello, L.; Musso, R.; Cillino, G.; Pitrone, M.; Pizzolanti, G.; Coppola, A.; Arancio, W.; Di Cara, G.; et al. Structure, Function, and Mechanism of Thioredoxin Proteins. Antioxid. Redox Signal. 2010, 13, 1205–1216. [Google Scholar] [CrossRef]
  19. Montrichard, F.; Alkhalfioui, F.; Yano, H.; Vensel, W.H.; Hurkman, W.J.; Buchanan, B.B. Thioredoxin targets in plants: The first 30 years. J. Proteom. 2009, 72, 452–474. [Google Scholar] [CrossRef]
  20. Balmer, Y.; Vensel, W.H.; Tanaka, C.K.; Hurkman, W.J.; Gelhaye, E.; Rouhier, N.; Jacquot, J.-P.; Manieri, W.; Schürmann, P.; Droux, M.; et al. Thioredoxin links redox to the regulation of fundamental processes of plant mitochondria. Proc. Natl. Acad. Sci. USA 2004, 101, 2642–2647. [Google Scholar] [CrossRef]
  21. Broin, M.; Rey, P. Potato Plants Lacking the CDSP32 Plastidic Thioredoxin Exhibit Overoxidation of the BAS1 2-Cysteine Peroxiredoxin and Increased Lipid Peroxidation in Thylakoids under Photooxidative Stress. Plant Physiol. 2003, 132, 1335–1343. [Google Scholar] [CrossRef]
  22. Laloi, C.; Mestres-Ortega, D.; Marco, Y.; Meyer, Y.; Reichheld, J.-P. The Arabidopsis Cytosolic Thioredoxin h5 Gene Induction by Oxidative Stress and Its W-Box-Mediated Response to Pathogen Elicitor. Plant Physiol. 2004, 134, 1006–1016. [Google Scholar] [CrossRef] [Green Version]
  23. Zhang, C.-J.; Zhao, B.-C.; Ge, W.-N.; Zhang, Y.-F.; Song, Y.; Sun, D.-Y.; Guo, Y. An Apoplastic H-Type Thioredoxin Is Involved in the Stress Response through Regulation of the Apoplastic Reactive Oxygen Species in Rice. Plant Physiol. 2011, 157, 1884–1899. [Google Scholar] [CrossRef]
  24. Lakhssassi, N.; Doblas, V.G.; Rosado, A.; del Valle, A.E.; Posé, D.; Jimenez, A.J.; Castillo, A.G.; Valpuesta, V.; Borsani, O.; Botella, M.A. The Arabidopsis tetratricopeptide thioredoxin-like Gene Family Is Required for Osmotic Stress Tolerance and Male Sporogenesis. Plant Physiol. 2012, 158, 1252–1266. [Google Scholar] [CrossRef]
  25. Åslund, F.; Beckwith, J. Bridge over Troubled Waters: Sensing Stress by Disulfide Bond Formation. Cell 1999, 96, 751–753. [Google Scholar] [CrossRef]
  26. Martí, M.C.; Florez-Sarasa, I.; Camejo, D.; Ribas-Carbó, M.; Lázaro, J.J.; Sevilla, F.; Jiménez, A. Response of mitochondrial thioredoxin PsTrxo1, antioxidant enzymes, and respiration to salinity in pea (Pisum sativum L.) leaves. J. Exp. Bot. 2011, 62, 3863–3874. [Google Scholar] [CrossRef]
  27. Sun, L.; Ren, H.; Liu, R.; Li, B.; Wu, T.; Sun, F.; Liu, H.; Wang, X.; Dong, H. An h-Type Thioredoxin Functions in Tobacco Defense Responses to Two Species of Viruses and an Abiotic Oxidative Stress. Mol. Plant-Microbe Interact. 2010, 23, 1470–1485. [Google Scholar] [CrossRef]
  28. Xie, G.; Kato, H.; Sasaki, K.; Imai, R. A cold-induced thioredoxin h of rice, OsTrx23, negatively regulates kinase activities of OsMPK3 and OsMPK6 in vitro. FEBS Lett. 2009, 583, 2734–2738. [Google Scholar] [CrossRef]
  29. Rorat, T.; Havaux, M.; Irzykowski, W.; Cuiné, S.; Becuwe, N.; Rey, P. PSII-S gene expression, photosynthetic activity and abundance of plastid thioredoxin-related and lipid-associated proteins during chilling stress in Solanum species differing in freezing resistance. Physiol. Plant. 2001, 113, 72–78. [Google Scholar] [CrossRef]
  30. Park, J.H.; Lee, E.S.; Chae, H.B.; Paeng, S.K.; Wi, S.D.; Bae, S.B.; Thi Phan, K.A.; Lee, S.Y. Disulfide reductase activity of thi-oredoxin-h2 imparts cold tolerance in Arabidopsis. Biochem. Biophys. Res. Commun. 2021, 568, 124–130. [Google Scholar] [PubMed]
  31. Lee, E.; Park, J.; Wi, S.; Chae, H.; Paeng, S.; Bae, S.; Phan, K.; Lee, S. Arabidopsis Disulfide Reductase, Trx-h2, Functions as an RNA Chaperone under Cold Stress. Appl. Sci. 2021, 11, 6865. [Google Scholar] [CrossRef]
  32. Lee, E.S.; Park, J.H.; Wi, S.D.; Kang, C.H.; Chi, Y.H.; Chae, H.B.; Paeng, S.K.; Ji, M.G.; Kim, W.-Y.; Kim, M.G.; et al. Redox-dependent structural switch and CBF activation confer freezing tolerance in plants. Nat. Plants 2021, 7, 914–922. [Google Scholar] [CrossRef]
  33. Kopeć, P.; Rapacz, M.; Arora, R. Post-translational activation of CBF for inducing freezing tolerance. Trends Plant Sci. 2022, 27, 415–417. [Google Scholar] [CrossRef] [PubMed]
  34. Wu, F.; Li, Q.; Yan, H.; Zhang, D.; Jiang, G.; Jiang, Y.; Duan, X. Characteristics of Three Thioredoxin Genes and Their Role in Chilling Tolerance of Harvested Banana Fruit. Int. J. Mol. Sci. 2016, 17, 1526. [Google Scholar] [CrossRef] [PubMed]
  35. Menges, M.; Hennig, L.; Gruissem, W.; Murray, J.A.H. Cell Cycle-regulated Gene Expression in Arabidopsis. J. Biol. Chem. 2002, 277, 41987–42002. [Google Scholar] [CrossRef] [PubMed]
  36. Bréhélin, C.; Mouaheb, N.; Verdoucq, L.; Lancelin, J.-M.; Meyer, Y. Characterization of Determinants for the Specificity of Arabidopsis Thioredoxins h in Yeast Complementation. J. Biol. Chem. 2000, 275, 31641–31647. [Google Scholar] [CrossRef]
  37. Park, S.K.; Jung, Y.J.; Lee, J.R.; Lee, Y.M.; Jang, H.H.; Lee, S.S.; Park, J.H.; Kim, S.Y.; Moon, J.C.; Lee, S.Y.; et al. Heat-Shock and Redox-Dependent Functional Switching of an h-Type Arabidopsis Thioredoxin from a Disulfide Reductase to a Molecular Chaperone. Plant Physiol. 2009, 150, 552–561. [Google Scholar] [CrossRef]
  38. Sweat, T.A.; Wolpert, T.J. Thioredoxin h5 Is Required for Victorin Sensitivity Mediated by a CC-NBS-LRR Gene in Arabidopsis. Plant Cell 2007, 19, 673–687. [Google Scholar] [CrossRef]
  39. Tada, Y.; Spoel, S.H.; Pajerowska-Mukhtar, K.; Mou, Z.; Song, J.; Wang, C.; Zuo, J.; Dong, X. Plant immunity requires con-formational changes of NPR1 via S-nitrosylation and thioredoxins. Science 2008, 321, 952–956. [Google Scholar] [CrossRef]
  40. Dos Santos, C.V.; Rey, P. Plant thioredoxins are key actors in the oxidative stress response. Trends Plant Sci. 2006, 11, 329–334. [Google Scholar] [CrossRef]
  41. Chae, H.B.; Moon, J.C.; Shin, M.R.; Chi, Y.H.; Jung, Y.J.; Lee, S.Y.; Nawkar, G.M.; Jung, H.S.; Hyun, J.K.; Kim, W.Y.; et al. Thioredoxin Reductase Type C (NTRC) Orchestrates Enhanced Thermotolerance to Arabidopsis by Its Redox-Dependent Holdase Chaperone Function. Mol. Plant 2013, 6, 323–336. [Google Scholar] [CrossRef]
  42. Belin, C.; Bashandy, T.; Cela, J.; Delorme-Hinoux, V.; Riondet, C.; Reichheld, J.P. A comprehensive study of thiol reduction gene expression under stress conditions in Arabidopsis thaliana. Plant Cell Environ. 2014, 38, 299–314. [Google Scholar] [CrossRef] [PubMed]
  43. Sevilla, F.; Camejo, D.; Ortiz-Espín, A.; Calderón, A.; Lázaro, J.J.; Jiménez, A. The thioredoxin/peroxiredoxin/sulfiredoxin system: Current overview on its redox function in plants and regulation by reactive oxygen and nitrogen species. J. Exp. Bot. 2015, 66, 2945–2955. [Google Scholar] [CrossRef]
  44. Da Fonseca-Pereira, P.; Daloso, D.M.; Gago, J.; Nunes-Nesi, A.; Araújo, W.L. On the role of the plant mitochondrial thioredoxin system during abiotic stress. Plant Signal. Behav. 2019, 14, 1592536. [Google Scholar] [CrossRef]
  45. Ben Saad, R.; Zouari, N.; Ben Ramdhan, W.; Azaza, J.; Meynard, D.; Guiderdoni, E.; Hassairi, A. Improved drought and salt stress tolerance in transgenic tobacco overexpressing a novel A20/AN1 zinc-finger “AlSAP” gene isolated from the halophyte grass Aeluropus littoralis. Plant Mol. Biol. 2010, 72, 171–190. [Google Scholar]
  46. Ben Saad, R.; Ben Romdhan, W.; Zouari, N.; Azaza, J.; Mieulet, D.; Verdeil, J.-L.; Guiderdoni, E.; Hassairi, A. Promoter of the AlSAP gene from the halophyte grass Aeluropus littoralis directs developmental-regulated, stress-inducible, and organ-specific gene expression in transgenic tobacco. Transgenic Res. 2010, 20, 1003–1018. [Google Scholar] [CrossRef]
  47. BEN Saad, R.; Fabre, D.; Mieulet, D.; Meynard, D.; Dingkuhn, M.; Al-Doss, A.; Guiderdoni, E.; Hassairi, A. Expression of the Aeluropus littoralis AlSAP gene in rice confers broad tolerance to abiotic stresses through maintenance of photosynthesis. Plant Cell Environ. 2011, 35, 626–643. [Google Scholar] [CrossRef]
  48. Ben Romdhane, W.; Ben Saad, R.; Meynard, D.; Verdeil, J.L.; Azaza, J.; Zouari, N.; Fki, L.; Guiderdoni, E.; Al-Doss, A.; Hassairi, A. Ectopic expression of Aeluropus littoralis plasma membrane protein gene AlTMP1 confers abiotic stress tolerance in trans-genic tobacco by improving water status and cation homeostasis. Int. J. Mol. Sci. 2017, 18, 692. [Google Scholar]
  49. Ben Saad, R.; Ben Romdhane, W.; Bouteraa, M.T.; Jrad, O.; Ben Hsouna, A. Lobularia maritima thioredoxin-h2 gene mitigates salt and osmotic stress damage in tobacco by modeling plant antioxidant system. Plant Growth Regul. 2022, 97, 101–115. [Google Scholar] [CrossRef]
  50. Ashraf, U.; Kanu, A.S.; Deng, Q.; Mo, Z.; Pan, S.; Tian, H.; Tang, X. Lead (Pb) Toxicity; Physio-Biochemical Mechanisms, Grain Yield, Quality, and Pb Distribution Proportions in Scented Rice. Front. Plant Sci. 2017, 8, 259. [Google Scholar] [CrossRef]
  51. Mittler, R. Oxidative stress, antioxidants and stress tolerance. Trends Plant Sci. 2002, 7, 405–410. [Google Scholar] [CrossRef] [PubMed]
  52. Duan, X.; Wang, Z.; Zhang, Y.; Li, H.; Yang, M.; Yin, H.; Cui, J.; Chai, H.; Gao, Y.; Hu, G.; et al. Overexpression of a thioredoxin-protein-encoding gene, MsTRX, from Medicago sativa enhances salt tolerance to transgenic tobacco. Agronomy 2022, 12, 1467. [Google Scholar]
  53. Moon, J.C.; Lee, S.; Shin, S.Y.; Chae, H.B.; Jung, Y.J.; Jung, H.S.; Lee, K.O.; Lee, J.R.; Lee, S.Y. Overexpression of Arabidopsis NADPH-dependent thioredoxin reductase C (AtNTRC) confers freezing and cold shock tolerance to plants. Biochem. Biophys. Res. Commun. 2015, 463, 1225–1229. [Google Scholar] [CrossRef] [PubMed]
  54. Tovar-Méndez, A.; Matamoros, M.A.; Bustos-Sanmamed, P.; Dietz, K.-J.; Cejudo, F.J.; Rouhier, N.; Sato, S.; Tabata, S.; Becana, M. Peroxiredoxins and NADPH-Dependent Thioredoxin Systems in the Model Legume Lotus japonicus. Plant Physiol. 2011, 156, 1535–1547. [Google Scholar] [CrossRef] [PubMed]
  55. Barajas-López, J.D.D.; Serrato, A.J.; Olmedilla, A.; Chueca, A.; Sahrawy, M. Localization in Roots and Flowers of Pea Chloroplastic Thioredoxin f and Thioredoxin m Proteins Reveals New Roles in Nonphotosynthetic Organs. Plant Physiol. 2007, 145, 946–960. [Google Scholar] [CrossRef] [Green Version]
  56. Renard, M.; Alkhalfioui, F.; Schmitt-Keichinger, C.; Ritzenthaler, C.; Montrichard, F. Identification and Characterization of Thioredoxin h Isoforms Differentially Expressed in Germinating Seeds of the Model Legume Medicago truncatula. Plant Physiol. 2011, 155, 1113–1126. [Google Scholar] [CrossRef]
  57. Lee, M.Y.; Shin, K.H.; Kim, Y.K.; Suh, J.Y.; Gu, Y.Y.; Kim, M.R.; Hur, Y.S.; Son, O.; Kim, J.S.; Song, E.; et al. Induction of thi-oredoxin is required for nodule development to reduce reactive oxygen species levels in soybean roots. Plant Physiol. 2005, 139, 1881–1889. [Google Scholar]
  58. Ji, M.G.; Park, H.J.; Cha, J.-Y.; Kim, J.A.; Shin, G.-I.; Jeong, S.Y.; Lee, E.S.; Yun, D.-J.; Lee, S.Y.; Kim, W.-Y. Expression of Arabidopsis thaliana Thioredoxin-h2 in Brassica napus enhances antioxidant defenses and improves salt tolerance. Plant Physiol. Biochem. 2020, 147, 313–321. [Google Scholar] [CrossRef]
  59. Elasad, M.; Ahmad, A.; Wang, H.; Ma, L.; Yu, S.; Wei, H. Overexpression of CDSP32 (GhTRX134) cotton gene enhances drought, salt, and oxidative stress tolerance in Arabidopsis. Plants 2020, 9, 1388. [Google Scholar]
  60. Noctor, G.; Mhamdi, A.; Foyer, C.H. The Roles of Reactive Oxygen Metabolism in Drought: Not So Cut and Dried. Plant Physiol. 2014, 164, 1636–1648. [Google Scholar] [CrossRef]
  61. Mittler, R.; Vanderauwera, S.; Gollery, M.; Van Breusegem, F. Reactive oxygen gene network of plants. Trends Plant Sci. 2004, 9, 490–498. [Google Scholar] [CrossRef] [PubMed]
  62. Li, Q.Y.; Niu, H.B.; Yin, J.; Shao, H.B.; Niu, J.S.; Ren, J.P.; Li, Y.C.; Wang, X. Transgenic barley with overexpressed PTrx in-creases aluminum resistance in roots during germination. J. Zhejiang Univ. Sci. B 2010, 11, 862–870. [Google Scholar]
  63. Park, S.; Lee, C.-M.; Doherty, C.J.; Gilmour, S.J.; Kim, Y.; Thomashow, M.F. Regulation of the Arabidopsis CBF regulon by a complex low-temperature regulatory network. Plant J. 2015, 82, 193–207. [Google Scholar] [CrossRef] [PubMed]
  64. Jia, Y.; Ding, Y.; Shi, Y.; Zhang, X.; Gong, Z.; Yang, S. The cbfs triple mutants reveal the essential functions of CBFs in cold acclimation and allow the definition of CBF regulons in Arabidopsis. New Phytol. 2016, 212, 345–353. [Google Scholar] [CrossRef] [PubMed]
  65. Zhao, C.; Zhang, Z.; Xie, S.; Si, T.; Li, Y.; Zhu, J.-K. Mutational Evidence for the Critical Role of CBF Transcription Factors in Cold Acclimation in Arabidopsis. Plant Physiol. 2016, 171, 2744–2759. [Google Scholar] [CrossRef]
  66. Vogel, J.T.; Zarka, D.G.; Van Buskirk, H.A.; Fowler, S.G.; Thomashow, M.F. Roles of the CBF2 and ZAT12 transcription factors in configuring the low temperature transcriptome of Arabidopsis. Plant J. 2005, 41, 195–211. [Google Scholar] [CrossRef]
  67. Narusaka, Y.; Nakashima, K.; Shinwari, Z.K.; Sakuma, Y.; Furihata, T.; Abe, H.; Narusaka, M.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Interaction between two cis-acting elements, ABRE and DRE, in ABA-dependent expression of Arabidopsis rd29A gene in response to dehydration and high-salinity stresses. Plant J. 2003, 34, 137–148. [Google Scholar] [CrossRef]
  68. Msanne, J.; Lin, J.S.; Stone, J.M.; Awada, T. Characterization of abiotic stress-responsive Arabidopsis thaliana RD29A and RD29B genes and evaluation of transgenes. Planta 2011, 234, 97–107. [Google Scholar] [CrossRef]
  69. Lee, S.Y.; Boon, N.J.; Webb, A.A.; Tanaka, R.J. Synergistic activation of RD29A via integration of salinity stress and abscisic acid in Arabidopsis thaliana. Plant Cell Physiol. 2016, 57, 2147–2160. [Google Scholar] [CrossRef]
  70. Hundertmark, M.; Hincha, D.K. LEA (Late Embryogenesis Abundant) proteins and their encoding genes in Arabidopsis thaliana. BMC Genom. 2008, 9, 118. [Google Scholar] [CrossRef]
  71. Jin, C.; Huang, X.-S.; Li, K.-Q.; Yin, H.; Li, L.-T.; Yao, Z.-H.; Zhang, S.-L. Overexpression of a bHLH1 Transcription Factor of Pyrus ussuriensis Confers Enhanced Cold Tolerance and Increases Expression of Stress-Responsive Genes. Front. Plant Sci. 2016, 7, 441. [Google Scholar] [CrossRef]
  72. Ben Romdhane, W.; Ben Saad, R.; Meynard, D.; Zouari, N.; Mahjoub, A.; Fki, L.; Guiderdoni, E.; Al-Doss, A.; Hassairi, A. Overexpression of AlTMP2 gene from the halophyte grass Aeluropus littoralis in transgenic tobacco enhances tolerance to different abiotic stresses by improving membrane stability and deregulating some stress-related genes. Protoplasma 2018, 255, 1161–1177. [Google Scholar]
  73. Zouari, N.; Ben Saad, R.; Legavre, T.; Azaza, J.; Sabau, X.; Jaoua, S.; Masmoudi, K.; Hassairi, A. Identification and sequencing of ESTs from the halophyte grass Aeluropus littoralis. Gene 2007, 404, 61–69. [Google Scholar] [CrossRef]
  74. Ben Saad, R.; Ben Halima, N.; Ghorbel, M.; Zouari, N.; Ben Romdhane, W.; Guiderdoni, E.; Al-Doss, A.; Hassairi, A. AlSRG1, a novel gene encoding an RRM-type RNA-binding protein (RBP) from Aeluropus littoralis, confers salt and drought tolerance in transgenic tobacco. Environ. Exp. Bot. 2018, 150, 25–36. [Google Scholar] [CrossRef]
  75. Livak, K.J.; Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCT Method. Methods 2001, 25, 402–408. [Google Scholar]
  76. Ben-Saad, R.; Ben Romdhane, W.; Zouari, N.; Azaza, J.; Mieulet, D.; Guiderdoni, E.; Ellouz, R.; Hassairi, A. Marker-free transgenic durum wheat cv. Karim expressing the AlSAP gene exhibits a high level of tolerance to salinity and dehydration stresses. Mol. Breed. 2011, 30, 521–533. [Google Scholar] [CrossRef]
  77. Arnon, D.I. Copper enzymes in isolated chloroplasts. Polyphenoloxidase in Beta vulgaris. Plant Physiol. 1949, 24, 1–15. [Google Scholar]
  78. Ben Hsouna, A.; Ghneim-Herrera, T.; Ben Romdhane, W.; Dabbous, A.; Ben Saad, R.; Brini, F.; Abdelly, C.; Ben Hamed, K. Early effects of salt stress on the physiological and oxidative status of the halophyte Lobularia maritima. Funct. Plant Biol. 2020, 47, 912. [Google Scholar] [CrossRef] [PubMed]
  79. Bouteraa, M.T.; Mishra, A.; Romdhane, W.B.; Hsouna, A.B.; Siddique, K.H.M.; Saad, R.B. Bio-stimulating effect of natural polysaccharides from Lobularia maritima on durum wheat Seedlings: Improved plant growth, salt stress tolerance by modulating biochemical responses and ion homeostasis. Plants 2022, 11, 1991. [Google Scholar]
  80. Chen, Q.; Heald, C.L.; Jimenez, J.L.; Canagaratna, M.R.; Zhang, Q.; He, L.; Huang, X.; Campuzano-Jost, P.; Palm, B.B.; Poulain, L.; et al. Elemental composition of organic aerosol: The gap between ambient and laboratory measurements. Geophys. Res. Lett. 2015, 42, 4182–4189. [Google Scholar] [CrossRef]
Figure 1. LmTrxh2 expression profiles and protein accumulation in response to cold stress (4 °C). (A) Pattern of expression of the LmTrxh2 gene and (B) the production of LmTrxh2 protein in the leaves and roots of L. maritima plants following the application of cold stress for 48 h. A western blot analysis of the total protein extracts (10 µg) was performed using anti-LmTrxh2 rabbit IgG (H + L) antibodies on leaves and roots tissues. The upper panel shows the detection of LmTrxh2, and the lower panel shows anti-β-actin for loading control. Vertical bars indicate the standard deviation calculated from three replicates. Values are mean ± SEM (n = 3). Different lowercases indicate a significant difference at p < 0.05.
Figure 1. LmTrxh2 expression profiles and protein accumulation in response to cold stress (4 °C). (A) Pattern of expression of the LmTrxh2 gene and (B) the production of LmTrxh2 protein in the leaves and roots of L. maritima plants following the application of cold stress for 48 h. A western blot analysis of the total protein extracts (10 µg) was performed using anti-LmTrxh2 rabbit IgG (H + L) antibodies on leaves and roots tissues. The upper panel shows the detection of LmTrxh2, and the lower panel shows anti-β-actin for loading control. Vertical bars indicate the standard deviation calculated from three replicates. Values are mean ± SEM (n = 3). Different lowercases indicate a significant difference at p < 0.05.
Ijms 24 03030 g001
Figure 2. Assessment of NT and LmTrxh2 transgenic lines performance at germination stage under cold stress conditions. (A) Photographs were taken two weeks after seed germination. (B) Seed germination rates were determined for LmTrxh2 overexpressors lines and NT plants under normal (25 °C) and cold stress (4 °C) conditions. The results presented are the means of three independent biological replicates, and a minimum of 30 seeds were counted for each experiment. (C) Comparison of the growth of transgenic and NT plants on plates. Tobacco seedlings were grown vertically for two weeks, and the fresh weight (D) and total chlorophyll content were measured (E) under normal and cold stress conditions. Data are expressed as the mean ± SEM (n = 3). Different lowercases indicate a significant difference at p < 0.05.
Figure 2. Assessment of NT and LmTrxh2 transgenic lines performance at germination stage under cold stress conditions. (A) Photographs were taken two weeks after seed germination. (B) Seed germination rates were determined for LmTrxh2 overexpressors lines and NT plants under normal (25 °C) and cold stress (4 °C) conditions. The results presented are the means of three independent biological replicates, and a minimum of 30 seeds were counted for each experiment. (C) Comparison of the growth of transgenic and NT plants on plates. Tobacco seedlings were grown vertically for two weeks, and the fresh weight (D) and total chlorophyll content were measured (E) under normal and cold stress conditions. Data are expressed as the mean ± SEM (n = 3). Different lowercases indicate a significant difference at p < 0.05.
Ijms 24 03030 g002
Figure 3. The LmTrxh2 overexpressing plant had increased cold tolerance. (A) The phenotypic appearance of 45-day old seedlings for NT and transgenic plants following the application of cold stress (4 °C) for 7 days, followed by recovery at 25 °C for 15 days. (B) Quantification of MDA, (C) H2O2, and (D) O2 accumulation in the leaves of NT and transgenic tobacco lines under control or subjected to cold stress (4 °C for 48 h). (E) The electrolyte leakage analysis of NT and LmTrxh2 transgenic lines under normal growth conditions and after cold treatment (4 °C for 48 h). The average of three independent experiments ± SEM is shown. Different lowercases indicate a significant difference at p < 0.05.
Figure 3. The LmTrxh2 overexpressing plant had increased cold tolerance. (A) The phenotypic appearance of 45-day old seedlings for NT and transgenic plants following the application of cold stress (4 °C) for 7 days, followed by recovery at 25 °C for 15 days. (B) Quantification of MDA, (C) H2O2, and (D) O2 accumulation in the leaves of NT and transgenic tobacco lines under control or subjected to cold stress (4 °C for 48 h). (E) The electrolyte leakage analysis of NT and LmTrxh2 transgenic lines under normal growth conditions and after cold treatment (4 °C for 48 h). The average of three independent experiments ± SEM is shown. Different lowercases indicate a significant difference at p < 0.05.
Ijms 24 03030 g003
Figure 4. Physiological indices of NT and LmTrxh2 transgenic plants under control and cold stress conditions. The leaves were sampled from plants grown under normal growth conditions and plants subjected to 48 h of cold stress (4 °C). (A) NBT and (B) DAB staining of leaves. (C) CAT activity. (D) SOD activity. (E) POD activity. Values are presented as means ± SEM values (n = 3). Different letters indicate a significant difference at p < 0.05.
Figure 4. Physiological indices of NT and LmTrxh2 transgenic plants under control and cold stress conditions. The leaves were sampled from plants grown under normal growth conditions and plants subjected to 48 h of cold stress (4 °C). (A) NBT and (B) DAB staining of leaves. (C) CAT activity. (D) SOD activity. (E) POD activity. Values are presented as means ± SEM values (n = 3). Different letters indicate a significant difference at p < 0.05.
Ijms 24 03030 g004
Figure 5. Relative expression of ROS-related (NtSOD and NtCAT1) and cold-responsive genes (NtERD10C, NtLEA5, NtDREB1A, NtCOR15A, NtCOR47, and NtKIN1) in NT and transgenic lines before and after 48 h of cold stress (4 °C). Values are presented as means ± SEM values (n = 3). Different letters indicate a significant difference at p < 0.05.
Figure 5. Relative expression of ROS-related (NtSOD and NtCAT1) and cold-responsive genes (NtERD10C, NtLEA5, NtDREB1A, NtCOR15A, NtCOR47, and NtKIN1) in NT and transgenic lines before and after 48 h of cold stress (4 °C). Values are presented as means ± SEM values (n = 3). Different letters indicate a significant difference at p < 0.05.
Ijms 24 03030 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ben Saad, R.; Ben Romdhane, W.; Baazaoui, N.; Bouteraa, M.T.; Chouaibi, Y.; Mnif, W.; Ben Hsouna, A.; Kačániová, M. Functional Characterization of Lobularia maritima LmTrxh2 Gene Involved in Cold Tolerance in Tobacco through Alleviation of ROS Damage to the Plasma Membrane. Int. J. Mol. Sci. 2023, 24, 3030. https://doi.org/10.3390/ijms24033030

AMA Style

Ben Saad R, Ben Romdhane W, Baazaoui N, Bouteraa MT, Chouaibi Y, Mnif W, Ben Hsouna A, Kačániová M. Functional Characterization of Lobularia maritima LmTrxh2 Gene Involved in Cold Tolerance in Tobacco through Alleviation of ROS Damage to the Plasma Membrane. International Journal of Molecular Sciences. 2023; 24(3):3030. https://doi.org/10.3390/ijms24033030

Chicago/Turabian Style

Ben Saad, Rania, Walid Ben Romdhane, Narjes Baazaoui, Mohamed Taieb Bouteraa, Yosra Chouaibi, Wissem Mnif, Anis Ben Hsouna, and Miroslava Kačániová. 2023. "Functional Characterization of Lobularia maritima LmTrxh2 Gene Involved in Cold Tolerance in Tobacco through Alleviation of ROS Damage to the Plasma Membrane" International Journal of Molecular Sciences 24, no. 3: 3030. https://doi.org/10.3390/ijms24033030

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