Next Article in Journal
Pectobacterium punjabense Causing Blackleg and Soft Rot of Potato: The First Report in the Russian Federation
Previous Article in Journal
Characterization of Lithuanian Tomato Varieties and Hybrids Using Phenotypic Traits and Molecular Markers
Previous Article in Special Issue
Can Photoselective Nets’ Influence Pollen Traits? A Case Study in ‘Matua’ and ‘Tomuri’ Kiwifruit Cultivars
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Functional Characterization of the Paeonia ostii P5CS Gene under Drought Stress

1
College of Horticulture and Landscape Architecture, Yangzhou University, Yangzhou 225009, China
2
Joint International Research Laboratory of Agriculture and Agri-Product Safety, The Ministry of Education of China, Yangzhou University, Yangzhou 225009, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Plants 2024, 13(15), 2145; https://doi.org/10.3390/plants13152145
Submission received: 9 July 2024 / Revised: 30 July 2024 / Accepted: 31 July 2024 / Published: 2 August 2024
(This article belongs to the Special Issue Growth, Development, and Stress Response of Horticulture Plants)

Abstract

:
With persistent elevation in global temperature, water scarcity becomes a major threat to plant growth and development, yield security, agricultural sustainability, and food production. Proline, as a key osmolyte and antioxidant, plays a critical role in regulating drought tolerance in plants, especially its key biosynthetic enzyme, delta-1-pyrroline-5-carboxylate synthase (P5CS), which always positively responds to drought stress. As an important woody oil crop, the expansion of Paeonia ostii cultivation needs to address the issue of plant drought tolerance. Here, we isolated a PoP5CS gene from P. ostii, with an open reading frame of 1842 bp encoding 613 amino acids. PoP5CS expression progressively increased in response to increasing drought stress, and it was localized in the cytoplasm. Silencing of PoP5CS in P. ostii reduced drought tolerance, accompanied by decreased proline content, elevated reactive oxygen species (ROS) accumulation, and increased relative electrical conductivity (REC) and malondialdehyde (MDA) levels. Conversely, overexpression of PoP5CS in Nicotiana tabacum plants enhanced drought resistance, manifested by increased proline levels, reduced ROS accumulation, and lower REC and MDA contents. This study isolates PoP5CS from P. ostii and validates its role in regulating drought tolerance, providing valuable genetic resources and theoretical insights for the development of drought-resistant P. ostii cultivars.

1. Introduction

In the face of a warming planet and increasingly unpredictable weather patterns, drought has emerged as a pressing global threat to agricultural productivity and ecosystem stability. As a primary environmental stressor, drought significantly impacts plant growth and development, often leading to reduced yield and even plant death [1]. To survive and thrive under these adverse conditions, plants have evolved intricate mechanisms of osmotic adjustment, which play a pivotal role in mitigating the deleterious effects of drought stress. Osmotic adjustment is a physiological response that involves the accumulation of compatible solutes, such as proline, glycine betaine, and soluble sugars, within plant cells to maintain turgor pressure and cellular homeostasis [2,3,4,5,6]. Recent studies have underscored the complexity and versatility of these mechanisms, revealing that different plant species and even cultivars within the same species exhibit distinct patterns of coping abilities and osmotic adjustment in response to drought [7,8].
Among the various adaptive mechanisms employed by plants to mitigate the detrimental effects of drought, the accumulation of proline stands out as a pivotal response. Proline, a cyclic subamino acid, has garnered considerable attention in recent years for its multifaceted roles in conferring drought tolerance to plants. Its accumulation under drought stress is not merely a passive consequence of cellular dehydration but an active adaptive strategy that serves to stabilize subcellular structures, maintain osmotic balance, and scavenge reactive oxygen species (ROS) generated during stress periods [9,10]. Proline biosynthesis is primarily mediated by the enzymes delta-1-pyrroline-5-carboxylate synthase (P5CS) and pyrroline-5-carboxylate reductase (P5CR), while its degradation is catalyzed by proline dehydrogenase (ProDH) and pyrroline-5-carboxylate dehydrogenase (P5CDH). This metabolic flux is tightly regulated in response to drought, with increased P5CS activity and suppressed ProDH activity leading to enhanced proline accumulation [11]. The functions of proline and the P5CS gene in plant responses to abiotic stresses have been explored in multiple species. As a key rate-limiting enzyme for proline biosynthesis, the P5CS gene was first isolated from Vigna aconifolia in 1992 by Hu et al. [12]. Since then, a series of plant P5CS genes have been cloned and their functions have received widespread attention. In the model plant Arabidopsis thaliana, P5CS protein is encoded by two differentially regulated genes, AtP5CS1 and AtP5CS2. The abundance of AtP5CS transcripts is tissue-specific and could be induced by drought, salinity, or ABA [13]. Zhang et al. [14] isolated the promoter of the A. thaliana P5CS gene and analyzed the effect of NaCl stress on P5CS promoter activity and found that NaCl stress induced an increase in P5CS promoter activity. Yamchi et al. [15] generated multiple A. thaliana P5CS gene overexpression lines and found that transgenic plants have higher proline expression levels and greater resistance to osmotic stress compared to non-transgenic plants. In addition to A. thaliana, the potential of the P5CS gene to respond to stress has also been reported in other species. For, example, in Oryza sativa, the OsP5CS1 gene’s expression was induced by a wide range of abiotic stresses (NaCl, drought, low temperature) and was commonly expressed in a variety of organs, whereas OsP5CS2 was specifically highly expressed in mature plants and induced by NaCl and mannitol [16]. Medicago truncatula MtP5CS2 has a positive role in osmotic regulation [17]. In Oriental hybrid lily Sorbonne, the expression of LhSorP5CS was upregulated by NaCl, mannitol, and ABA and accompanied by more proline accumulation [18]. Therefore, studying proline accumulation under drought stress and the function of the P5CS gene could provide valuable insights into how plants adapt to and survive under water deficit conditions.
Paeonia ostii, originating from China, is a perennial woody oil crop with ornamental, medicinal, and economic values [19]. Its seed oil boasts a high content of unsaturated fatty acids up to 92%, including 42% of α-linolenic acid, making it popular in the market due to its health benefits. Given the scarcity of arable land in China, P. ostii is predominantly cultivated in arid regions such as hills and slopes, where water scarcity significantly affects its growth, development, and yield [20]. For further expansion of P. ostii’s area for cultivation in arid and semi-arid regions, previous scholars have worked on finding ways to alleviate drought damage in P. ostii. Several cultivation practices including exogenous calcium treatment, silicon treatment, ferulic acid treatment, graphene oxide treatment, and fulvic acid treatment have been utilized to enhance the mitigation of this stress [19,21,22,23,24]. At the level of drought-tolerant gene identification, potential drought-tolerant structural genes (CCoAOMT, LACS4) and the transcription factor STAT were unearthed and were thought to be involved in drought tolerance in P. ostii [25,26,27]. However, the relationship between the P5CS gene and drought tolerance in P. ostii has not been reported. Here, a P5CS gene positively responding to drought stress was identified in P. ostii. Through gene cloning, expression pattern analysis, subcellular localization observation, and virus-induced PoP5CS gene silencing and overexpression experiments, we explored the positive role of the PoP5CS gene in regulating drought tolerance in P. ostii. This is the first characterization of the PoP5CS gene’s function in P. ostii drought regulation, which contributes to our understanding of the molecular mechanism of proline-regulated drought tolerance in P. ostii.

2. Results

2.1. PoP5CS Positively Responds to Drought Stress

Based on a drought-related transcriptome database of P. ostii [28], we worked to find drought-responsive genes that regulate drought tolerance in P. ostii. We focused on the key osmotic regulation-associated proline biosynthesis gene, P5CS. We found a total of three differential P5CS genes (Unigene0006017, Unigene0000731, and Unigene0008549) (Table S1). Among them, Unigene0008549 had the highest FPKM value and the highest differential fold of 4.04, suggesting its positive response to drought stress. Then, the full length of Unigene0008549 was obtained by PCR, and its coding sequence was 1842 bp in length, encoding 613 amino acids. To resolve the evolutionary relationship between Unigene0008549 and other homologs, we downloaded eight P5CS proteins from the NCBI database and applied them to an evolutionary tree analysis with Unigene0008549. As shown in Figure 1A, Unigene0008549 clustered with A. thaliana AtP5CS1 and AtP5CS2 and shared 57% homology with them, indicating that Unigene0008549 belonged to the P5CS family, therefore named as PoP5CS. Multiple sequence comparison showed that there existed a high sequence similarity between PoP5CS and P5CS family members (Figure 1B). Subsequently, the expression profiles of PoP5CS in P. ostii leaves sampled on Days 0, 4, 8, and 12 after drought treatment were scrupulously examined by quantitative real-time PCR (RT-qPCR) analysis. As shown in Figure 2, the expression level of PoP5CS showed a continuous increase, which was 1.82 times higher than that of Day 0 on Day 4, 20.61 times higher than that of Day 0 on Day 8, and reached an astonishing multiplicity of 260.37 on Day 12. The above results were consistent with its upregulated expression pattern in the transcriptome database, which was further evidence of the importance of PoP5CS for P. ostii to respond to drought stress. The above results revealed that PoP5CS positively responded to drought stress.

2.2. PoP5CS is Located in the Cytoplasm

In order to further reveal the expression characteristics of PoP5CS protein, we utilized Nicotiana benthamiana leaves for subcellular localization. We transiently transformed fusion vectors containing P5CS protein and enhanced green fluorescent protein (eGFP) into N. benthamiana leaves. The fusion vector expressed in the A. thaliana RPW8 2 gene (accession number: NM_114906.2) and mCherry fusion protein was used to indicate cytoplasmic localization signals. Following infiltration, fluorescent signals at 488 nm and 561 nm were elicited in N. benthamiana leaves to clarify the protein localization of PoP5CS. The emission of the eGFP signal, originating from the 35 S promoter-driven fusion construct of PoP5CS-eGFP, was observed within the cytoplasmic compartment of N. benthamiana leaves, where it co-localized with a cytoplasmic marker protein (Figure 3). This finding provided evidence that the subcellular localization of the PoP5CS protein occurs specifically within the cytoplasm of N. benthamiana leaves.

2.3. PoP5CS-Silenced P. ostii Show More Sensitivity to Drought

To investigate the function of PoP5CS in regulating drought tolerance, we obtained control plants (P. ostii transformed with pTRV), as well as PoP5CS-silenced P. ostii line 1 and line 2 based on the virus-induced gene silencing (VIGS) technique, and cultured them to the leaf expanding stage. Before drought treatment, there was no significant difference in plant phenotypes among these plants (Figure 4A). Prior to the drought treatment, we treated PoP5CS-silenced P. ostii line 2 with foliar proline spray for 3 days, and control plants (P. ostii transformed with pTRV) and PoP5CS-silenced P. ostii line 1 were sprayed with deionized water as controls. Next, the above P. ostii plants were subjected to drought treatment. DNA and total RNA were extracted from control plants (P. ostii transformed with pTRV) and PoP5CS-silenced P. ostii line 1 and line 2 on Day 15 after drought treatment for positive transgenic plant validation. As shown in Figure 4B, when DNA was amplified using P. ostii Ubiquitin-specific primers and pTRV1-specific primers, all experimental plants had a bright band, whereas when pTRV2-specific primers were used to amplify DNA, the pTRV band position differed in length from PoP5CS-silenced P. ostii line 1 and line 2. In addition, we quantified the expression level of PoP5CS in the above plants and found that the expression level of PoP5CS in PoP5CS-silenced P. ostii line 1 and line 2 decreased compared with pTRV by 35.45% and 30.23%. When they were exposed to drought treatment, the control plants (P. ostii transformed with pTRV) and proline-treated PoP5CS-silenced P. ostii line 2 maintained a normal growth condition and did not show a drought damage phenotype (Figure 4A). In contrast, PoP5CS-silenced P. ostii line 1 plants showed drooping stems and wilted leaves, which might be attributed to the suppressed expression of PoP5CS (Figure 4A). Then, several physiological experiments were applied to investigate the deeper response of PoP5CS to drought. As shown in Figure 5A,B, H2O2 accumulation identified by diaminobenzidine (DAB) staining and O2·− accumulation identified by nitrotetrazolium blue chloride (NBT) staining were used to assess the extent of ROS accumulation in P. ostii leaves. Among them, PoP5CS-silenced P. ostii line 1 plants showed large areas of brown and blue coloration in the leaves, implying a severe degree of drought damage. In contrast, control plants (P. ostii transformed with pTRV) and PoP5CS-silenced P. ostii line 2 plants showed little or almost no coloration in their leaves. Moreover, leaf water content, relative electrical conductivity (REC), and malondialdehyde (MDA) content all showed low levels under normal conditions (Figure 5C–E). When they were exposed to drought, PoP5CS-silenced P. ostii line 1 exhibited a more substantial decrease in water content, while control plants (P. ostii transformed with pTRV) and PoP5CS-silenced P. ostii line 2 showed a smaller decrease (Figure 5C). REC and MDA contents increased to a high extent in PoP5CS-silenced P. ostii line 1 and to a lesser extent in control plants (P. ostii transformed with pTRV) and PoP5CS-silenced P. ostii line 2 (Figure 5D,E). The above results suggest that PoP5CS played a positive role in regulating drought tolerance in P. ostii.

2.4. Overexpression of PoP5CS Increases Drought Tolerance in Nicotiana tabacum

To further confirm the function of PoP5CS in plant drought resistance, we generated two dependent PoP5CS transgenic N. tabacum lines, named line 1 and line 2. T2 generations of N. tabacum were cultivated for drought treatment. Under normal conditions, there was no significant difference between wild-type (WT) and PoP5CS transgenic N. tabacum lines (Figure 6A). Then, WT and PoP5CS transgenic N. tabacum lines were subjected to drought treatment. DNA and total RNA were extracted from WT and PoP5CS transgenic N. tabacum lines on Day 10 after drought treatment for positive transgenic plant validation. As shown in Figure 6B, when DNA was amplified using Hygromycin (hygromycin label region of pCAMBIA1301 vector)-specific primers and PoP5CS-specific primers, only a bright band existed in the lanes of PoP5CS transgenic N. tabacum lines. In addition, we quantified the expression level of PoP5CS in the above plants and found that the expression level of PoP5CS in PoP5CS transgenic N. tabacum lines were 29.59-fold (line 1) and 47.89-fold greater (line 2) when compared with WT (Figure 6C). After passing through 10 days of drought treatment, PoP5CS transgenic N. tabacum lines grew well in comparison to the severe drought damage condition of WT. The corresponding physiological data also matched the phenotypic observations (Figure 7). Among them, ROS were over-accumulated in WT and the water content of WT was significantly reduced compared with PoP5CS transgenic N. tabacum lines, while REC and MDA contents were significantly increased in the WT (Figure 7A–E). The above results confirmed the results that PoP5CS could increase drought tolerance in plants.

2.5. PoP5CS Regulates Drought Tolerance by Promoting Proline Accumulation

To deeper investigate the effects of PoP5CS, plant proline contents were determined in the control plants (P. ostii transformed with pTRV), PoP5CS-silenced P. ostii, WT, and PoP5CS transgenic N. tabacum (Figure 5F and Figure 7F). Under normal conditions, the proline content of the control plants (P. ostii transformed with pTRV) was significantly higher than the PoP5CS-silenced P. ostii lines with an average 1.95-fold difference. At the same time, the proline content of the WT was slightly lower than the PoP5CS transgenic N. tabacum. After drought treatment, proline contents were substantially increased in all of the above plants. In VIGS assay, the proline contents in the control plants (P. ostii transformed with pTRV) and PoP5CS-silenced P. ostii line 2 were much higher than PoP5CS-silenced P. ostii line 1. In the overexpression assay, the proline contents in PoP5CS transgenic N. tabacum lines were much higher than the WT. Obviously, the proline content was positively correlated with the drought tolerance of plants. These findings suggest that PoP5CS increased the drought tolerance of plants by promoting proline accumulation.

3. Discussion

3.1. PoP5CS is a Member of the P5CS Family

When plants are subjected to stress, a significant accumulation of proline occurs within their bodies, thereby enhancing their adaptive capacity to stress. As a crucial osmotic regulatory substance, proline plays a pivotal role in lowering the water potential of plant cells, strengthening their water absorption and retention capabilities, and thus safeguarding cellular osmotic balance and subcellular structural stability [29,30]. P5CS genes, a key role in plant proline biosynthesis, have been reported in a variety of plant species, such as A. thaliana, Cajanus cajan, N. benthamiana, and Oryza sativa [31,32,33,34,35]. In this study, we isolated the PoP5CS gene from P. ostii. The coding sequence of the PoP5CS gene was 1842 bp in length, encoding 613 amino acids. Liu and Wang [36] cloned the AmP5CS gene from Avicennia marina, and the coding sequence of AmP5CS was 2148 bp (encoded 715 amino acid protein). Yang et al. [37] isolated the SpP5CS gene from Stipa purpurea. SpP5CS contained a coding sequence of 2196 bp which encoded 731 amino acids. Then, a phylogenetic tree clustered PoP5CS into the branch containing A. thaliana AtP5CS1 and AtP5CS2. PoP5CS, Theobroma cacao TcP5CS, and Gossypium hirsutum GhP5CS were involved in multiple sequence comparison, and they had high sequence similarity. These indicated that PoP5CS is a typical P5CS family member.

3.2. PoP5CS Exhibits a Positive Response to Drought Stress Conditions

Exploring the expression patterns of genes under different environmental conditions can help to reveal their potential roles in cellular processes and involved metabolic pathways. P5CS genes play an important role in plant responses to a variety of abiotic stresses, such as drought, high temperature, and low temperature, largely determined by their expression response to different environments [32,38]. In O. sativa, salt-tolerant variants of plants exhibited an elevated level of OsP5CS mRNA transcripts under conditions of high salinity, alongside an increase in proline content, compared to their expression under standard conditions [31]. In reed canary grass, expression pattern analysis indicated an upregulation of B231P5CS transcripts in leaves subsequent to salt treatment (200 mM NaCl), with peak transcript abundance observed 12 h post-exposure, marking a 14.69-fold increase compared to untreated control plants [39]. In Carica papaya, an investigation into the expression dynamics of CpP5CS2 under both low (7 °C)- and high (35 °C)-temperature stresses was conducted. The findings revealed that both stress conditions triggered an upregulation of CpP5CS2 expression during storage, with this enhancement preceding the accumulation of proline [38]. In S. purpurea, the expression of the SpP5CS gene demonstrated a comprehensive responsiveness to a spectrum of stress conditions, encompassing cold, salt, and PEG stresses. Notably, a markedly significant upregulation of SpP5CS gene expression was observed specifically during soil drought conditions and upon subsequent rehydration treatment [37]. Here, we analyzed the dynamic expression pattern of PoP5CS in P. ostii plants subjected to drought stress. After 4 days of drought treatment, the expression of PoP5CS reached 1.86 times that under untreated conditions, increased to 20.61-fold at Day 8, and remarkably surged to 260.37-fold at 12 days of drought treatment. Combining the above results, PoP5CS strongly responded to drought stress, and this response may bring about a large accumulation of proline.

3.3. PoP5CS Improves Plant Drought Resistance by Accumulating Proline Content

Plants exhibit resilience against drought stress through a diverse array of intricate mechanisms, with cellular osmoregulation emerging as a pivotal mode of physiological adaptation. Proline, a pivotal osmolyte and antioxidant, assumes a paramount role in safeguarding the maintenance of cellular osmotic equilibrium and modulating the delicate balance of ROS within plant cells, thereby contributing significantly to the plant’s drought tolerance capabilities [9]. With this in mind, we attempted to validate the drought tolerance function of the PoP5CS gene in plants as well as to analyze the regulatory mechanisms by which they modulate drought tolerance. Through the VIGS technology and overexpression system, we obtained PoP5CS-silenced P. ostii and PoP5CS-overexpressing N. tabacum and observed their phenotypic responses to drought stress as well as changes in key physiological indicators. After drought treatment, the drought tolerance of these transgenic plants showed a significantly positive correlation with the expression levels of PoP5CS. The indirect assessment of cellular membrane integrity can be quantified through the REC of cellular leakages [40]. PoP5CS-silenced P. ostii had higher REC compared with the control, while exogenous spraying of proline reduced the REC of PoP5CS-silenced P. ostii. Similarly, PoP5CS-overexpressing N. tabacum demonstrated lower REC than the WT. MDA represents the degree of membrane lipid peroxidation, which indirectly reflects the degree of damage to the cell membrane system [41]. MDA accumulation in PoP5CS-silenced P. ostii leaves was significantly higher than the control, while MDA in proline-treated PoP5CS-silenced P. ostii leaves was significantly lower than that in deionized water-treated PoP5CS-silenced P. ostii. Concordantly, PoP5CS-overexpressing N. tabacum accumulated considerably less MDA relative to the WT. In addition, we assessed the abundance of ROS accumulation in transgenic plants by DAB and NBT staining and found that ROS accumulation in PoP5CS-silenced P. ostii was significantly higher than that of the control, whereas exogenous spraying of proline reduced the level of ROS accumulation in PoP5CS-silenced P. ostii. Likewise, ROS barely accumulated in PoP5CS-overexpressing N. tabacum. In a previous study, Chen et al. [42] treated Phaseolus vulgaris PvP5CS transgenic A. thaliana with NaCl, and found that the REC of transgenic PvP5CS A. thaliana was significantly lower than control plants. In Saccharum officinarum, the P5CS gene is a potential drought tolerance target gene, and overexpression of SoP5CS resulted in lower MDA content compared with the WT [43]. In our study, we also found that plant drought tolerance, as well as indicators of drought damage, showed a high degree of consistency with proline content, and that high proline accumulation ensured that plants maintained normal growth after drought stress, such as proline-treated PoP5CS-silenced P. ostii and PoP5CS-overexpressing N. tabacum. Similar results have been identified in multiple species, such as S. purpurea, Lilium regale, and N. benthamiana [33,37,44]. Our study validated, for the first time, the positive function of PoP5CS, a key gene for proline biosynthesis, in mediating proline-mediated drought tolerance in P. ostii, providing insights into the improvement of drought tolerance in P. ostii.

4. Materials and Methods

4.1. Plant Materials and Drought Treatment

Potting soil mixed with loam, peat, and perlite (1:1:1) was used to cultivate one-year-old plants of P. ostii. After three-day continuous watering, P. ostii plants were subjected to natural drought in a greenhouse without water supplementation. P. ostii leaves were collected on Days 0, 4, 8, and 12 after drought treatment. For the VIGS assay, transgenic P. ostii plants at the leaf expanding stage were subjected to a 15-day-drought treatment. For the heterologous overexpression assay, the T2 generation of transgenic N. tabacum was subjected to a 10-day drought treatment. For the subcellular localization assay, N. benthamiana seeds were sown in the above potting soil and incubated under long sunlight (25 °C 16 h light/22 °C 8 h dark).

4.2. Gene Cloning, Multiple Sequence Alignment, and Phylogenetic Analysis

We first extracted total RNA from P. ostii leaves and then synthesized complementary DNA (cDNA) using EasyScript RT/RI Enzyme Mix (TransGen, Beijing, China). For PoP5CS gene cloning, specific primers (Table S2) were designed based on reference gene sequences derived from the P. ostii transcriptome database (SRA: SRP161474). The 2 × Phanta Max Master DNA (Vazyme, Nanjing, China) was applied to amplify the PoP5CS gene, and then, purified products were ligated into the 5 × TA/Blunt-Zero Cloning Mix (Vazyme, Nanjing, China) and amplified by Escherichia coli. The sequence was confirmed by Tsingke (Nanjing, China). For phylogenetic analysis, the amino acid sequence of PoP5CS was used to search its homologs in other species based on the E-value algorithm in the National Center for Biotechnology Information (NCBI) database. Also, the P5CS proteins from the model plant A. thaliana and several woody plants (Populus trichocarpa, Malus domestica, Pyrus × bretschneideri, and Vitis vinifera) for which a full genome sequence was available were also included in the phylogenetic tree analysis. MEGA7.0 was used for protein alignment by Clustal W, and a neighbor-joining phylogenetic tree was constructed by the bootstrap method with 1000 replications, p-distance, and pairwise deletion. Sequence comparison was performed by DNAMAN 6.0.

4.3. RT-qPCR Analysis

Total RNA from P. ostii leaves (including temporally drought-treated leaves and leaves with gene silencing) and N. tabacum leaves with gene overexpression was extracted. Then, cDNA for RT-qPCR analysis was synthesized by 5 × EasyScript All-in-One SuperMix (TransGen, Beijing, China). NovoStart SYBR qPCR Super Mix (Novoprotein, Suzhou, China) was employed in conjunction with a BIO-RAD CFX Connect Optics Module (Bio-Rad, Hercules, CA, USA) to detect gene fluorescence signals, and the comprehensive experimental procedures can be found in Yu et al. [45]. The 2−△△Ct method was employed to derive the gene relative expression level. For normalization, P. ostii Ubiquitin (JN699053) and N. tabacum tubulin alpha chain (TUBA, NP_001312557.2) served as internal reference controls. The primers used here can be found in Table S2.

4.4. Subcellular Localization

The PoP5CS coding sequence was amplified by gene-specific primers (forward 5′-CGGGGATCCTCTAGAGTCGACATGGCTCTTTATGATACTTTGTTCAGC-3′, reverse 5′-CACCATGGTACTAGTGTCGACCGACAACAGTGGGAGATCCTTG-3′) and then fused into a GFP tagged p35S::GFP vector; 100 ng p35S::PoP5CS-GFP, empty p35S::GFP vectors, and pCAMBIA1300-35S-NES-mCherry vectors (expressing cytoplasmic localization signals directed RPW8 2 protein) were transformed into Agrobacterium tumefaciens, respectively. The A. tumefaciens bacterial solution was resuspended in infiltration buffer (10 mM MES, 10 mM MgCl2, and 0.2 mM acetosyringone) and adjusted to a standard concentration (OD = 0.7). Then, the A. tumefaciens bacterial solution was used to infect the leaves of approximately 30-day-old N. benthamiana. Two days after infection, the localization regions of PoP5CS were observed by confocal laser microscopy (Nikon C2-ER, Tokyo, Japan).

4.5. VIGS Assay

The VIGS approach utilizing the pTRV vectors was employed to transiently silence the PoP5CS gene in P. ostii. In the TRV-based VIGS system, pTRV1 and pTRV2 are the two key viral vector components, each of which has a different function and works together to achieve the gene silencing effect [46]. pTRV2 is responsible for silencing specific genes, and pTRV1 is responsible for the replication and systematic spread of plasmids in plants. The non-conserved fragment of PoP5CS (from 20 bp to 338 bp) was fused into a pTRV2 vector with gene-specific primers (forward 5′-AAGGTTACCGAATTCTCTAGAGTATGGATTAGCTTTGAATGTGGATC-3′, reverse 5′-CGTGAGCTCGGTACCGGATCCTGTTCAGCCAGCTTGATGTGAC-3′); 100 ng pTRV2-PoP5CS, pTRV2 and pTRV1 vectors were transformed into A. tumefaciens, respectively. A. tumefaciens cultures were cultured overnight in a shaker at 28 °C to achieve an OD of 1.5. The resuspended A. tumefaciens bacterial solution mixtures (pTRV2-PoP5CS mixed with pTRV1, pTRV2 mixed with pTRV1) were kept in the dark for 1 h. P. ostii plants at the dormant stage with 2~3 buds were used as infestation materials. After pruning to create wounds, the roots were washed 2~3 times using deionized water. Then, the roots were immersed in the above bacterial solution, and a vacuum desiccator was employed to create negative pressure for 30 min, facilitating the penetration of the bacterial solution into the plants. Subsequent to being washed twice with deionized water, the plants were repotted into potting soil, with each experimental group comprising 15 individual plants. After a period of 30 days under cultivation, the plants had reached the leaf expansion stage. For the drought tolerance assays, the control plants (P. ostii transformed with pTRV) and PoP5CS-silenced P. ostii line 1 plants were treated with deionized water, and PoP5CS-silenced P. ostii line 2 was treated with proline for 3 days to test the function of proline on P. ostii drought resistance following approximately 15 days of drought treatment. The primers used here can be found in Table S2.

4.6. Heterologous Overexpression Assay

The PoP5CS coding sequence was amplified by gene-specific primers (forward 5′-CAGGTCGACTCTAGAGGATCCATGGCTCTTTATGATACTTTGTTCAGC-3′, reverse 5′-CGATCGGGGAAATTCGAGCTCCGACAACAGTGGGAGATCCTTG-3′) and then fused into a pCAMBIA1301 vector. The pCAMBIA1301-PoP5CS vector was introduced for N. tabacum transformation using the leaf disc method via T-DNA insertion as previously described [47]. Briefly, the pCAMBIA1301-PoP5CS vector was transformed into A. tumefaciens; the A. tumefaciens bacterial solution was resuspended in infiltration buffer (10 mM MES, 10 mM MgCl2, and 0.2 mM acetosyringone) and adjusted to a standard concentration (OD = 0.3). Then, we took leaves from N. tabacum sterile seedlings, infiltrated the leaves using the above mentioned bacterial solution, filtered out the solution, and co-cultivated it for 3 days in darkness. The leaves were then transferred to resistant shoot screening and differentiation medium for selection culture. The differentiated adventitious shoots were transferred to rooting screening medium for rooting culture to obtain the T0 generation. Positive PoP5CS transgenic N. tabacum was screened by antibiotic labeling, and two independent lines of PoP5CS transgenic N. tabacum (T2 generation) were used to verify their function in drought tolerance. Positive PoP5CS transgenic N. tabacum seedlings rooted in the medium were transplanted into potting soil to carry out conventional greenhouse-based cultivation. After approximately 3 months, PoP5CS transgenic N. tabacum plants with similar phenotypes were subjected to a 10-day natural drought treatment as above to verify their drought resistance function. The primers used here can be found in Table S2.

4.7. Physiological Index Measurement

DNA was first extracted from PoP5CS transgenic P. ostii leaves and PoP5CS transgenic N. tabacum leaves. Then, PCR and RT-qPCR were applied for positive plant validation (Table S2). Both leaves from PoP5CS transgenic P. ostii and PoP5CS transgenic N. tabacum before and after drought treatment were used for physiological index measurement. The specific experimental procedures of water content determination, REC measurement, MDA content measurement, and DAB and NBT staining were followed as previously described [48]. The proline content was measured using a reagent kit (Comin Biotechnology Co., Ltd., Suzhou, China). Briefly, 0.1 g of PoP5CS transgenic P. ostii leaves and PoP5CS transgenic N. tabacum leaves were extracted by 1 mL sulfosalicylic acid solution. After homogenization in an ice bath, the mixture was subjected to oscillation extraction at 90 °C for 10 min, followed by centrifugation at 10,000× g and 25 °C for 10 min, with the supernatant being collected. To 0.5 mL of the supernatant, 0.5 mL of glacial acetic acid and 0.5 mL of reaction mixture (containing ninhydrin, glacial acetic acid, and concentrated phosphoric acid solution) were added. After thorough mixing, the solution was kept in a boiling water bath for 30 min. After cooling, 1 mL of toluene was added to the test tube, followed by shaking for 30 s. The absorbance at a wavelength of 520 nm was measured to calculate the proline content.

4.8. Statistical Analysis

All data consisted of at least three biological replicates and error bars indicate standard error. The significant differences were analyzed by one-way ANOVA (p < 0.05) or by Student’s t-test (* p < 0.05; ** p < 0.01; *** p < 0.001).

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/plants13152145/s1, Table S1: Transcriptome data of differential P5CS genes; Table S2: Primers used in this study.

Author Contributions

Conceptualization, J.T.; methodology, J.T. and D.Z.; software, Y.L., H.A. and Z.C.; data curation, Y.L., H.A. and Z.C.; writing—original draft preparation, Y.L.; writing—review and editing, Y.L., J.T. and D.Z.; supervision, J.T. and D.Z.; project administration, J.T. and D.Z.; funding acquisition, J.T. and D.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Long-term scientific research base for conservation and innovative utilization of Peony germplasm resources in Jiangsu Province (LYKJ[2021]01), the Central finance forestry science and technology promotion demonstration fund project (Su[2023]TG05), the Agriculture Science and Technology Innovation Foundation of Jiangsu Province (CX[22]3186), the High-Level Talent Support Program of Yangzhou University, and National Forest and Grass Science and Technology Innovation and Development Research Project (2023132012).

Data Availability Statement

Data are contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Chaves, M.M.; Maroco, J.P.; Pereira, J.S. Understanding plant responses to drought—From genes to the whole plant. Funct. Plant Biol. 2003, 30, 239–264. [Google Scholar] [CrossRef] [PubMed]
  2. Quan, R.D.; Shang, M.; Zhang, H.; Zhao, Y.X.; Zhang, J.R. Engineering of enhanced glycine betaine synthesis improves drought tolerance in maize. Plant Biotechnol. J. 2004, 2, 477–486. [Google Scholar] [CrossRef] [PubMed]
  3. Ashraf, M.; Foolad, M.R. Roles of glycine betaine and proline in improving plant abiotic stress resistance. Environ. Exp. Bot. 2007, 59, 206–216. [Google Scholar] [CrossRef]
  4. Farooq, M.; Wahid, A.; Kobayashi, N.; Fujita, D.; Basra, S.M.A. Plant drought stress: Effects, mechanisms and management. Agron. Sustain. Dev. 2009, 29, 185–212. [Google Scholar] [CrossRef]
  5. Ahammed, G.J.; Li, X.; Wan, H.J.; Zhou, G.Z.; Cheng, Y. SlWRKY81 reduces drought tolerance by attenuating proline biosynthesis in tomato. Sci. Hortic. 2020, 270, 109444. [Google Scholar] [CrossRef]
  6. O’Brien, M.J.; Valtat, A.; Abiven, S.; Studer, M.S.; Ong, R.; Schmid, B. The role of soluble sugars during drought in tropical tree seedlings with contrasting tolerances. J. Plant Ecol. 2020, 13, 389–397. [Google Scholar] [CrossRef]
  7. Wu, G.Q.; Zhang, L.N.; Wang, Y.Y. Response of growth and antioxidant enzymes to osmotic stress in two different wheat (Triticum aestivum L.) cultivars seedlings. Plant Soil. Environ. 2012, 58, 534–539. [Google Scholar] [CrossRef]
  8. De Caroli, M.; Rampino, P.; Curci, L.M.; Pecatelli, G.; Carrozzo, S.; Piro, G. CiXTH29 and CiLEA4 role in water stress tolerance in Cichorium intybus varieties. Biology 2023, 12, 444. [Google Scholar] [CrossRef]
  9. Verslues, P.E.; Sharma, S. Proline metabolism and its implications for plant-environment interaction. Arab. Book 2010, 8, e0140. [Google Scholar] [CrossRef]
  10. Szabados, L.; Savouré, A. Proline: A multifunctional amino acid. Trends Plant Sci. 2010, 15, 89–97. [Google Scholar] [CrossRef]
  11. Alvarez, M.E.; Savouré, A.; Szabados, L. Proline metabolism as regulatory hub. 2022. Trends Plant Sci. 2022, 27, 39–55. [Google Scholar] [CrossRef] [PubMed]
  12. Hu, C.A.; Delauney, A.J.; Verma, D.P.S. A bifunctional enzyme (delta 1-pyrroline-5-carboxylate synthetase) catalyzes the first two steps in proline biosynthesis in plants. Proc. Natl. Acad. Sci. USA 1992, 89, 9354–9358. [Google Scholar] [CrossRef] [PubMed]
  13. Strizhov, N.; Abraham, E.; Okresz, L.; Blickling, S.; Zilberstein, A.; Schell, J.; Koncz, C.; Szabados, L. Differential expression of two P5CS genes controlling proline accumulation during salt-stress requires ABA and is regulated by ABA1, AB/1 and AXR2 in Arabidopsis. Plant J. 1997, 12, 557–569. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, C.S.; Lu, Q.; Verma, D.P.S. Characterization of Δ1-pyrroline-5-carboxylate synthetase gene promoter in transgenic Arabidopsis thaliana subjected to water stress. Plant Sci. 1997, 129, 81–89. [Google Scholar] [CrossRef]
  15. Yamchi, A.; Jazii, F.R.; Mousavi, A.; Karkhane, A.A. Proline accumulation in transgenic tobacco as a result of expression of Arabidopsis Δ1-pyrroline-5-carboxylate synthetase (P5CS) during osmotic stress. J. Plant Biochem. Biot. 2007, 16, 9–15. [Google Scholar] [CrossRef]
  16. Hur, J.; Jung, K.H.; Lee, C.H.; An, G. Stress-inducible OsP5CS2 gene is essential for salt and coldtolerance in rice. Plant Sci. 2004, 167, 417–426. [Google Scholar] [CrossRef]
  17. Verdoy, D.; Coba De La Peña, T.; Redondo, F.J.; Lucas, M.M.; Pueyo, J.J. Transgenic Medicago truncatula plants that accumulate proline display nitrogen-fixing activity with enhanced tolerance to osmotic stress. Plant Cell Environ. 2006, 29, 1913–1923. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, L.; Guo, Z.H.; Zhang, Y.B.; Wang, Y.J.; Yang, G.; Yang, L.; Wang, R.Y.; Xie, Z.K. Characterization of LhsorP5CS, a gene catalyzing proline synthesis in oriental hybrid Lily sorbonne: Molecularmodelling and expression analysis. Bot. Stud. 2017, 58, 10. [Google Scholar] [CrossRef]
  19. Shi, X.L.; Xue, X.; Xu, H.M.; Yang, Y.Q.; Kuang, Z.X.; Hou, X.G. Amelioration of salt-induced damage on Paeonia ostii 'Fengdan' by exogenous application of silicon. Agronomy 2023, 13, 1349. [Google Scholar] [CrossRef]
  20. Guo, L.L.; Li, Y.Y.; Wei, Z.Z.; Wang, C.; Hou, X.G. Reference genes selection of Paeonia ostii ‘Fengdan’ under osmotic stresses and hormone treatments by RT-qPCR. Mol. Biol. Rep. 2023, 50, 133–143. [Google Scholar] [CrossRef]
  21. Zhang, X.Y.; Fang, Z.W.; Liu, H.N.; Zhao, D.Q.; Tao, J. Exogenous calcium-induced physiological and biochemical changes in tree peony (Paeonia section Moutan DC.) under drought stress. Photosynthetica 2019, 57, 904–911. [Google Scholar] [CrossRef]
  22. Zhao, D.Q.; Fang, Z.W.; Tang, Y.H.; Tao, J. Graphene oxide as an effective soil water retention agent can confer drought stress tolerance to Paeonia ostii without toxicity. Environ. Sci. Technol. 2020, 54, 8269–8279. [Google Scholar] [CrossRef]
  23. Fang, Z.W.; Wang, X.X.; Zhang, X.Y.; Zhao, D.Q.; Tao, J. Effects of fulvic acid on the photosynthetic and physiological characteristics of Paeonia ostii under drought stress. Plant Signal. Behav. 2020, 15, e1774714. [Google Scholar] [CrossRef] [PubMed]
  24. Fang, Z.W.; Zhang, X.Y.; Tao, J.; Zhao, D.Q. Ameliorative effect of ferulic acid on Paeonia ostii under drought stress. Bull. Bot. Res. 2020, 40, 353–359. [Google Scholar]
  25. Zhao, D.Q.; Luan, Y.T.; Shi, W.B.; Zhang, X.Y.; Meng, J.S.; Tao, J. A Paeonia ostii caffeoyl-CoA O-methyltransferase confers drought stress tolerance by promoting lignin synthesis and ROS scavenging. Plant Sci. 2021, 303, 110765. [Google Scholar] [CrossRef]
  26. Zhang, H.Y.; Zhang, S.; Li, M.; Wang, J.; Wu, T. The PoLACS4 gene may participate in drought stress resistance in tree peony (Paeonia ostii ‘Feng Dan Bai’). Genes 2022, 13, 1591. [Google Scholar] [CrossRef] [PubMed]
  27. Guo, L.L.; Shen, J.J.; Zhang, C.J.; Guo, Q.; Liang, H.Y.; Hou, X.G. Characterization and bioinformatics analysis of ptc-miR396g-5p in response to drought stress of Paeonia ostii. Non-Coding RNA Res. 2022, 7, 150–158. [Google Scholar] [CrossRef] [PubMed]
  28. Zhao, D.Q.; Zhang, X.Y.; Fang, Z.W.; Wu, Y.Q.; Tao, J. Physiological and transcriptomic analysis of tree peony (Paeonia section Moutan DC.) in response to drought stress. Forests 2019, 10, 135. [Google Scholar] [CrossRef]
  29. Trovato, M.; Mattioli, R.; Costantino, P. Multiple roles of proline in plant stress tolerance and development. Rend. Lincel-Sci. Fis. 2008, 19, 325–346. [Google Scholar] [CrossRef]
  30. Mattioli, R.; Costantino, P.; Trovato, M. Proline accumulation in plants: Not only stress. Plant Signal. Behav. 2009, 4, 1016–1018. [Google Scholar] [CrossRef]
  31. Igarashi, Y.; Yoshiba, Y.; Sanada, Y.; Yamaguchi-Shinozaki, K.; Wada, K.; Shinozaki, K. Characterization of the gene for DELTA-1-pyrroline-5-carboxylate synthetase and correlation between the expression of the gene and salt tolerance in Oryza sativa L. Plant Mol. Biol. 1997, 33, 857–865. [Google Scholar] [CrossRef]
  32. Su, J.; Wu, R. Stress-inducible synthesis of proline in transgenic rice confers faster growth under stress conditions than that with constitutive synthesis. Plant Sci. 2004, 166, 941–948. [Google Scholar] [CrossRef]
  33. Ku, H.M.; Hu, C.C.; Chang, H.J.; Lin, Y.T.; Jan, F.J.; Chen, C.T. Analysis by virus induced gene silencing of the expression of two proline biosynthetic pathway genes in Nicotiana benthamiana under stress conditions. Plant Physiol. Bioch. 2011, 49, 1147–1154. [Google Scholar] [CrossRef] [PubMed]
  34. Surekha, C.; Kumari, K.N.; Aruna, L.V.; Suneetha, G.; Arundhati, A.; Kishor, P.B.K. Expression of the Vigna aconitifolia P5CSF129A gene in transgenic pigeonpea enhances proline accumulation and salt tolerance. Plant Cell Tiss. Org. 2014, 116, 27–36. [Google Scholar] [CrossRef]
  35. Guo, C.J.; Zhang, T.Y.; Leng, Q.Q.; Zhou, X.; Zhong, J.L.; Liu, J.L. Dynamic Arabidopsis P5CS filament facilitates substrate channelling. Nat. Plants 2024, 10, 880–889. [Google Scholar] [CrossRef]
  36. Liu, J.; Wang, Y.S. Proline metabolism and molecular cloning of AmP5CS in the mangrove Avicennia marina under heat stress. Ecotoxicology 2020, 29, 698–706. [Google Scholar] [CrossRef] [PubMed]
  37. Yang, D.N.; Ni, R.Z.; Yang, S.A.; Pu, Y.A.; Qian, M.; Yang, Y.Q.; Yang, Y.P. Functional characterization of the Stipa purpurea P5CS gene under drought stress conditions. Int. J. Mol. Sci. 2021, 22, 9599. [Google Scholar] [CrossRef]
  38. Zhu, X.Y.; Li, X.P.; Zou, Y.; Chen, W.X.; Lu, W.J. Cloning, characterization and expression analysis of Δ1-pyrroline-5-carboxylate synthetase (P5CS) gene in harvested papaya (Carica papaya) fruit under temperature stress. Food Res. Int. 2012, 49, 272–279. [Google Scholar] [CrossRef]
  39. Cong, L.L.; Zhang, X.Q.; Yang, F.Y.; Liu, S.J.; Zhang, Y.W. Isolation of the P5CS gene from reed canary grass and its expression under salt stress. Genet. Mol. Res. 2014, 13, 9122–9133. [Google Scholar] [CrossRef]
  40. Bajji, M.; Kinet, J.M.; Lutts, S. The use of the electrolyte leakage method for assessing cell membrane stability as a water stresstolerance test in durum wheat. Plant Growth Regul. 2002, 36, 61–70. [Google Scholar] [CrossRef]
  41. Morliere, P.; Moysan, A.; Santus, R.; Huppe, G.; Maziere, J.C.; Dubertret, L. UVA-induced lipid peroxidation in cultured human fibroblasts. BBA-Biomembr. 1991, 1084, 261–268. [Google Scholar] [CrossRef] [PubMed]
  42. Chen, J.B.; Yang, J.W.; Zhang, Z.Y.; Feng, X.F.; Wang, S.M. Two P5CS genes from common bean exhibiting different tolerance to salt stress in transgenic Arabidopsis. J. Genet. 2013, 92, 461–469. [Google Scholar] [CrossRef]
  43. Li, J.; Phan, T.T.; Li, Y.R.; Xing, Y.X.; Yang, L.T. Isolation, transformation and overexpression of sugarcane SoP5CS gene for drought tolerance improvement. Sugar Tech 2018, 20, 464–473. [Google Scholar] [CrossRef]
  44. Wei, C.; Cui, Q.; Zhang, X.Q.; Zhao, Y.Q.; Jia, G.X. Three P5CS genes including a novel one from Lilium regale play distinct roles in osmotic, drought and salt stress tolerance. J. Plant Biol. 2016, 59, 456–466. [Google Scholar] [CrossRef]
  45. Yu, Q.X.; Du, H.; Huang, Y.Y.; Lei, X.; Wu, X.T.; Jiang, J.Y.; Huang, W.; Ge, L.F. KINASE-INDUCIBLE DOMAIN INTERACTING 8 regulates helical pod morphology in Medicago truncatula. Plant Physiol. 2024, 195, 2016–2031. [Google Scholar] [CrossRef]
  46. Liu, Y.L.; Schiff, M.; Marathe, R.; Dinesh-Kumar, S.P. Tobacco Rar1, EDS1 and NPR1/NIM1 like genes are required for N-mediated resistance to tobacco mosaic virus. Plant J. 2002, 30, 415–429. [Google Scholar] [CrossRef] [PubMed]
  47. Sunilkumar, G.; Vijayachandra, K.; Veluthambi, K. Preincubation of cut tobacco leaf explants promotes Agrobacterium-mediated transformation by increasing vir gene induction. Plant Sci. 1999, 141, 51–58. [Google Scholar] [CrossRef]
  48. Huang, X.S.; Wang, W.; Zhang, Q.; Liu, J.H. A basic helix-loop-helix transcription factor, PtrbHLH, of Poncirus trifoliata confers cold tolerance and modulates peroxidase-mediated scavenging of hydrogen peroxide. Plant Physiol. 2013, 162, 1178–1194. [Google Scholar] [CrossRef]
Figure 1. Sequences analysis of PoP5CS. (A) A phylogenetic tree constructed using PoP5CS and P5CS proteins from other plant species. TcP5CS (Theobroma cacao, XP_007026912.2), DzP5CS-like (Durio zibethinus, XP_022766207.1), HtP5CS2 (Hibiscus trionum, GMJ03143.1), GhP5CS (Gossypium hirsutum, XP_040936303.1), GrP5CS2 (Gossypium raimondii, XP_012458611.1), AtP5CS1 (Arabidopsis thaliana, NP_001189714.1), AtP5CS2 (A. thaliana, NP_191120.2), MaP5CS (Melia azedarach, KAJ4704590.1), PtP5CS (Populus trichocarpa, XP_002315202.1), MdP5CS-like (Malus domestica, XP_008387828.2), VvP5CS (Vitis vinifera, XP_010658318.1), PbP5CS (Pyrus × bretschneideri, XP_009343959.2). PoP5CS is marked with a red pentagram. The neighbor-joining tree was generated by MEGA7.0 using a p-distance model with 1000 bootstrap replicates. The tree is a bootstrap consensus tree, and bootstrap values are shown at the nodes. (B) Multiple sequence alignment of PoP5CS and homologous proteins from other species. PoP5CS is marked with a red pentagram. The colors in the sequence alignment mean different homology levels. Orange means homology level ≥ 33%, green means homology level ≥ 50%, and blue means homology level = 100%.
Figure 1. Sequences analysis of PoP5CS. (A) A phylogenetic tree constructed using PoP5CS and P5CS proteins from other plant species. TcP5CS (Theobroma cacao, XP_007026912.2), DzP5CS-like (Durio zibethinus, XP_022766207.1), HtP5CS2 (Hibiscus trionum, GMJ03143.1), GhP5CS (Gossypium hirsutum, XP_040936303.1), GrP5CS2 (Gossypium raimondii, XP_012458611.1), AtP5CS1 (Arabidopsis thaliana, NP_001189714.1), AtP5CS2 (A. thaliana, NP_191120.2), MaP5CS (Melia azedarach, KAJ4704590.1), PtP5CS (Populus trichocarpa, XP_002315202.1), MdP5CS-like (Malus domestica, XP_008387828.2), VvP5CS (Vitis vinifera, XP_010658318.1), PbP5CS (Pyrus × bretschneideri, XP_009343959.2). PoP5CS is marked with a red pentagram. The neighbor-joining tree was generated by MEGA7.0 using a p-distance model with 1000 bootstrap replicates. The tree is a bootstrap consensus tree, and bootstrap values are shown at the nodes. (B) Multiple sequence alignment of PoP5CS and homologous proteins from other species. PoP5CS is marked with a red pentagram. The colors in the sequence alignment mean different homology levels. Orange means homology level ≥ 33%, green means homology level ≥ 50%, and blue means homology level = 100%.
Plants 13 02145 g001
Figure 2. RT-qPCR analysis of the expression levels of PoP5CS under drought stress conditions on Days 0, 4, 8, and 12. PoP5CS expression on Day 0 is normalized to 1. Data represent the mean ± SD of three biological replicates, and letters indicate significant differences by one-way ANOVA (p < 0.05).
Figure 2. RT-qPCR analysis of the expression levels of PoP5CS under drought stress conditions on Days 0, 4, 8, and 12. PoP5CS expression on Day 0 is normalized to 1. Data represent the mean ± SD of three biological replicates, and letters indicate significant differences by one-way ANOVA (p < 0.05).
Plants 13 02145 g002
Figure 3. Subcellular localization of PoP5CS in Nicotiana benthamiana leaves. (a,e) The green fluorescence signals of the 35S:: eGFP expression vector and 35S:: PoP5CS-eGFP expression vector at 488 nm; (b,f) the red fluorescence signals of the RPW8 2-mCherry expression vector at 561 nm; (c,g) bright field images; (d) an overlaid image of (ac); (h) an overlaid image of (eg).
Figure 3. Subcellular localization of PoP5CS in Nicotiana benthamiana leaves. (a,e) The green fluorescence signals of the 35S:: eGFP expression vector and 35S:: PoP5CS-eGFP expression vector at 488 nm; (b,f) the red fluorescence signals of the RPW8 2-mCherry expression vector at 561 nm; (c,g) bright field images; (d) an overlaid image of (ac); (h) an overlaid image of (eg).
Plants 13 02145 g003
Figure 4. Phenotypic observation and positive plant identification of PoP5CS-silenced Paeonia ostii. (A) Phenotypic observation of PoP5CS-silenced P. ostii when PoP5CS was silenced. Bar = 10 cm. (B) PCR validation of PoP5CS-silenced P. ostii after 15 days of drought treatment. (C) RT-qPCR validation of PoP5CS-silenced P. ostii after 15 days of drought treatment. Data represent the mean ± SD of three replicates, and statistical significance was determined by Student’s t-test (** p < 0.01).
Figure 4. Phenotypic observation and positive plant identification of PoP5CS-silenced Paeonia ostii. (A) Phenotypic observation of PoP5CS-silenced P. ostii when PoP5CS was silenced. Bar = 10 cm. (B) PCR validation of PoP5CS-silenced P. ostii after 15 days of drought treatment. (C) RT-qPCR validation of PoP5CS-silenced P. ostii after 15 days of drought treatment. Data represent the mean ± SD of three replicates, and statistical significance was determined by Student’s t-test (** p < 0.01).
Plants 13 02145 g004
Figure 5. Physiological index measurement of PoP5CS-silenced Paeonia ostii. (A) H2O2 accumulation condition of PoP5CS-silenced P. ostii leaves after 15 days of drought treatment. Bar = 1 cm. (B) O2 accumulation condition of PoP5CS-silenced P. ostii leaves after 15 days of drought treatment. Bar = 1 cm. (C) Leaf water content when PoP5CS was silenced. (D) REC when PoP5CS was silenced. REC, relative electrical conductivity. (E) MDA content when PoP5CS was silenced. MDA, malondialdehyde. (F) Proline content when PoP5CS was silenced. Data represent the mean ± SD of three replicates, and letters indicated significant differences by one-way ANOVA (p < 0.05).
Figure 5. Physiological index measurement of PoP5CS-silenced Paeonia ostii. (A) H2O2 accumulation condition of PoP5CS-silenced P. ostii leaves after 15 days of drought treatment. Bar = 1 cm. (B) O2 accumulation condition of PoP5CS-silenced P. ostii leaves after 15 days of drought treatment. Bar = 1 cm. (C) Leaf water content when PoP5CS was silenced. (D) REC when PoP5CS was silenced. REC, relative electrical conductivity. (E) MDA content when PoP5CS was silenced. MDA, malondialdehyde. (F) Proline content when PoP5CS was silenced. Data represent the mean ± SD of three replicates, and letters indicated significant differences by one-way ANOVA (p < 0.05).
Plants 13 02145 g005
Figure 6. Phenotypic observation and positive plant identification of PoP5CS-overexpressing Nicotiana tabacum. (A) Phenotypic observation of PoP5CS-overexpressing N. tabacum when PoP5CS was overexpressed. Bar = 10 cm. (B) PCR validation of PoP5CS-overexpressing N. tabacum after 10 days of drought treatment. NtTUBA, specific primers in N. tabacum tubulin alpha chain coding sequence; Hyg, specific primers in the hygromycin label region of pCAMBIA1301 vector; PoP5CS, specific primers in PoP5CS coding sequence. (C) RT-qPCR validation of PoP5CS-overexpressing N. tabacum after 10 days of drought treatment. Data represent the mean ± SD of three replicates, and statistical significance was determined by Student’s t-test (*** p < 0.001).
Figure 6. Phenotypic observation and positive plant identification of PoP5CS-overexpressing Nicotiana tabacum. (A) Phenotypic observation of PoP5CS-overexpressing N. tabacum when PoP5CS was overexpressed. Bar = 10 cm. (B) PCR validation of PoP5CS-overexpressing N. tabacum after 10 days of drought treatment. NtTUBA, specific primers in N. tabacum tubulin alpha chain coding sequence; Hyg, specific primers in the hygromycin label region of pCAMBIA1301 vector; PoP5CS, specific primers in PoP5CS coding sequence. (C) RT-qPCR validation of PoP5CS-overexpressing N. tabacum after 10 days of drought treatment. Data represent the mean ± SD of three replicates, and statistical significance was determined by Student’s t-test (*** p < 0.001).
Plants 13 02145 g006
Figure 7. Physiological index measurement of PoP5CS-overexpressing Nicotiana tabacum. (A) H2O2 accumulation condition of PoP5CS-overexpressing N. tabacum leaves after 10 days of drought treatment. Bar = 3 cm. (B) O2 accumulation condition of PoP5CS-overexpressing N. tabacum leaves after 10 days of drought treatment. Bar = 3 cm. (C) Leaf water content when PoP5CS was overexpressed. (D) REC when PoP5CS was overexpressed. REC, relative electrical conductivity. (E) MDA content when PoP5CS was overexpressed. MDA, malondialdehyde. (F) Proline content when PoP5CS was overexpressed. Data represent the mean ± SD of three replicates, and letters indicated significant differences by one-way ANOVA (p < 0.05).
Figure 7. Physiological index measurement of PoP5CS-overexpressing Nicotiana tabacum. (A) H2O2 accumulation condition of PoP5CS-overexpressing N. tabacum leaves after 10 days of drought treatment. Bar = 3 cm. (B) O2 accumulation condition of PoP5CS-overexpressing N. tabacum leaves after 10 days of drought treatment. Bar = 3 cm. (C) Leaf water content when PoP5CS was overexpressed. (D) REC when PoP5CS was overexpressed. REC, relative electrical conductivity. (E) MDA content when PoP5CS was overexpressed. MDA, malondialdehyde. (F) Proline content when PoP5CS was overexpressed. Data represent the mean ± SD of three replicates, and letters indicated significant differences by one-way ANOVA (p < 0.05).
Plants 13 02145 g007
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

Luan, Y.; An, H.; Chen, Z.; Zhao, D.; Tao, J. Functional Characterization of the Paeonia ostii P5CS Gene under Drought Stress. Plants 2024, 13, 2145. https://doi.org/10.3390/plants13152145

AMA Style

Luan Y, An H, Chen Z, Zhao D, Tao J. Functional Characterization of the Paeonia ostii P5CS Gene under Drought Stress. Plants. 2024; 13(15):2145. https://doi.org/10.3390/plants13152145

Chicago/Turabian Style

Luan, Yuting, Honglei An, Zijie Chen, Daqiu Zhao, and Jun Tao. 2024. "Functional Characterization of the Paeonia ostii P5CS Gene under Drought Stress" Plants 13, no. 15: 2145. https://doi.org/10.3390/plants13152145

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