Next Article in Journal
Restraint Stress-Induced Neutrophil Inflammation Contributes to Concurrent Gastrointestinal Injury in Mice
Previous Article in Journal
Epigenetic Modulation of GPER Expression in Gastric and Colonic Smooth Muscle of Male and Female Non-Obese Diabetic (NOD) Mice: Insights into H3K4me3 and H3K27ac Modifications
Previous Article in Special Issue
Evolution of the WRKY Family in Angiosperms and Functional Diversity under Environmental Stress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Physiological and Proteomic Responses of the Tetraploid Robinia pseudoacacia L. to High CO2 Levels

1
College of Forestry and Grassland, Jilin Agriculture University, Changchun 130118, China
2
College of Life Science, Northeast Forestry University, Harbin 150040, China
3
Jilin Provincial Key Laboratory of Tree and Grass Genetics and Breeding, Changchun 130118, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2024, 25(10), 5262; https://doi.org/10.3390/ijms25105262
Submission received: 28 March 2024 / Revised: 6 May 2024 / Accepted: 9 May 2024 / Published: 11 May 2024

Abstract

:
The increase in atmospheric CO2 concentration is a significant factor in triggering global warming. CO2 is essential for plant photosynthesis, but excessive CO2 can negatively impact photosynthesis and its associated physiological and biochemical processes. The tetraploid Robinia pseudoacacia L., a superior and improved variety, exhibits high tolerance to abiotic stress. In this study, we investigated the physiological and proteomic response mechanisms of the tetraploid R. pseudoacacia under high CO2 treatment. The results of our physiological and biochemical analyses revealed that a 5% high concentration of CO2 hindered the growth and development of the tetraploid R. pseudoacacia and caused severe damage to the leaves. Additionally, it significantly reduced photosynthetic parameters such as Pn, Gs, Tr, and Ci, as well as respiration. The levels of chlorophyll (Chl a and b) and the fluorescent parameters of chlorophyll (Fm, Fv/Fm, qP, and ETR) also significantly decreased. Conversely, the levels of ROS (H2O2 and O2·−) were significantly increased, while the activities of antioxidant enzymes (SOD, CAT, GR, and APX) were significantly decreased. Furthermore, high CO2 induced stomatal closure by promoting the accumulation of ROS and NO in guard cells. Through a proteomic analysis, we identified a total of 1652 DAPs after high CO2 treatment. GO functional annotation revealed that these DAPs were mainly associated with redox activity, catalytic activity, and ion binding. KEGG analysis showed an enrichment of DAPs in metabolic pathways, secondary metabolite biosynthesis, amino acid biosynthesis, and photosynthetic pathways. Overall, our study provides valuable insights into the adaptation mechanisms of the tetraploid R. pseudoacacia to high CO2.

Graphical Abstract

1. Introduction

CO2 is a crucial substrate for plant photosynthesis. The levels of atmospheric CO2 have a significant impact on plant growth, development, and biomass [1]. Since the industrial era, atmospheric CO2 concentrations have increased by more than 40%, and the current ambient CO2 concentrations exceed 417 ppm [2]. The rise in CO2 concentration has resulted in significant alterations in global temperatures, intensifying the greenhouse effect and exposing plants to elevated CO2 levels, higher temperatures, and drought. Consequently, this poses a considerable challenge to plant growth and reproduction. The impact of CO2 enrichment on growth and yield for C3 plants can vary and may be influenced by species-specific factors [3].
Previous studies have consistently shown that higher levels of CO2 have a significant impact on the physiological functioning of plants [4,5,6]. On the one hand, elevated CO2 directly affects plant photosynthesis. Increased atmospheric CO2 concentrations lead to a higher number of leaves, longer branches, and greater biomass while also causing a decrease in stomatal density. Additionally, it can result in visible symptoms such as wilting and drooping leaves [3,7]. According to previous studies, the net photosynthetic rate and water use efficiency (WUE) of A. marina and R. stylosa increased when exposed to a slight elevation in CO2 concentration. These findings suggest that the promotion of photosynthesis was facilitated by the elevated CO2 concentration [7]. Furthermore, it has been observed that rising CO2 concentrations contribute to an increase in phytoplankton growth rate and lipid productivity [8]. Another study found that corn exhibits higher biomass and leaf area, as well as enhanced starch synthesis, when exposed to elevated CO2 concentrations [9]. The Rubisco content and light-saturated photosynthetic rate in plant leaves were observed to be significantly reduced in a high CO2 environment [10]. Similarly, when Pinus sylvestris was exposed to this environment for an extended period, both the stomatal conductance and total stomatal number showed a significant decrease [11]. On the other hand, in Allium sativum L., the elevated CO2 atmosphere had a contrasting effect, significantly enhancing the activity of pyruvate decarboxylase (PDC) and alcohol dehydrogenase (ADH). This, in turn, caused a substantial increase in the content of alcohols, aldehydes, and phenols, resulting in toxic effects [12]. On the contrary, research has demonstrated that high CO2 levels can also actually mitigate ozone-induced oxidative damage in wheat [8]. Additionally, the impact of CO2 concentrations varies depending on the plant species. For example, in the case of reticulated melons, low CO2 has been found to increase both the initial and maximum photosynthetic efficiency. However, when exposed to high CO2, the maximum photosynthetic efficiency of reticulated melons actually declines [13].
On the other hand, previous studies have demonstrated that CO2-induced stomatal closure is dependent on Ca2+ and that elevated CO2 levels stimulate an increase in free Ca2+ concentration in defense cells [14,15,16,17]. Stomata, which are small pores on the leaf surface, play a crucial role in regulating the uptake of carbon dioxide for photosynthesis and the loss of water through transpiration [18]. The movement of stomata is influenced by various environmental factors, including CO2 concentration. Stomatal opening is promoted by low concentrations of CO2, while stomatal closure is induced by high concentrations of CO2. In the guard cells of Arabidopsis thaliana, the carbonic anhydrases βCA1 and βCA4 are involved in mediating CO2-controlled stomatal movement and development [19]. Additionally, BIG proteins play a critical role in inhibiting stomatal opening and promoting stomatal closure in response to CO2 [20]. Reactive oxygen species (ROS) and nitric oxide (NO) are important signaling molecules that regulate stomatal movement in plants [21]. It has been shown that CO2-mediated stomatal closure requires the generation of ROS, which has an important role in the regulation of stomatal aperture [22,23]. NO plays a role in both abscisic acid (ABA)- and Ca2+-mediated stomatal closure and is located downstream of ROS, an important participant in the stomatal closure process [24,25]. In A. thaliana, ROS are mainly produced by NADPH oxidase and peroxidase at the plasma membrane [26,27], while NO is mainly produced by nitrate reductase (NR) and NOA1 and induces stomatal closure [28].
Plant respiration is a significant metabolic process on a global scale, accounting for approximately 40–60% of the rate of carbon assimilation [29]. Mitochondria, which are essential organelles involved in cellular material and energy metabolism, possess two respiratory pathways on their inner mitochondrial membrane [30]. There are two pathways involved in respiration: the cytochrome pathway, which is found in nearly all organisms, and the alternate respiration pathway. The alternate respiration pathway is activated when the cytochrome pathway experiences stress conditions like low temperature, drought, ozone, and high salinity soils [31]. The impact of CO2 on mitochondrial respiration is of great significance in the investigation of plant respiratory metabolism. Previous studies have demonstrated that there exists an alternate respiration pathway, which is insensitive to cyanide (CN) but sensitive to SHAM and is facilitated by alternate oxidase (AOX) [32,33,34]. Additionally, pyruvate, a crucial metabolite in organisms, has been found to have a notable enhancing effect on the activity of AOX [35,36,37].
The respiration rate serves as a primary indicator of plant respiratory metabolism. By examining the respiration rate of plants, we can directly observe the impact of high CO2 concentration on plant respiratory metabolism. However, the effect of elevated CO2 concentration on the plant respiration rate varies depending on the specific plants and environmental conditions. It has been discovered that elevated CO2 concentrations resulted in an 8.4% decrease in the daily respiration rate of wheat and sunflower, as well as a 16.2% decrease in the respiration rate during darkness. This could be attributed to the reduction in leaf nitrogen content and the downregulation of photosynthesis caused by elevated CO2 concentrations, which subsequently leads to a decrease in plant respiration rate [38]. However, previous studies have demonstrated that elevated CO2 concentrations have the potential to enhance respiratory substrate utilization, upregulate respiratory genes, and increase the number of mitochondria. As a result, this leads to an overall increase in the rate of plant respiration [39,40].
The tetraploid R. pseudoacacia L. is a deciduous ornamental tree species belonging to the genus R. pseudoacacia in the family Leguminosae. It has four sets of chromosomes (4n = 44) and was created through the application of colchicine to the diploid R. pseudoacacia (2n = 22) of a homologous species. This tetraploid species was introduced to China from South Korea [41]. Compared to the diploid R. pseudoacacia, the tetraploid R. pseudoacacia is considered an excellent variety due to its higher yield, faster growth rate, larger leaves, and numerous outstanding characteristics such as drought resistance, saline resistance, low-temperature resistance, and heat resistance [42,43,44,45]. The tetraploid R. pseudoacacia exhibits a robust root system, strong adaptability to various environments, and a high capacity for photosynthetic carbon sequestration [46]. In light of prevalent environmental pollution and global climate change, it is imperative to investigate the effects of high CO2 levels on the physiological processes of the tetraploid R. pseudoacacia.
Plants display distinct physiological and biochemical traits when subjected to prolonged exposure to elevated CO2 concentrations. This study focuses on the tetraploid R. pseudoacacia as the experimental material to examine the stomatal movement pattern and physiological alterations in photosynthesis and respiration in response to high CO2 conditions. Additionally, a proteomic approach was employed to identify crucial regulatory networks and proteins associated with the response of the tetraploid R. pseudoacacia to high CO2 treatment. The objective of this research is to gain insights into the mechanisms by which plants respond to high CO2 environments.

2. Results

2.1. Response of Morphological and Gas Exchange Parameters of Tetraploid R. pseudoacacia to High CO2

Under normal CO2 conditions, the growth of the tetraploid R. pseudoacacia was uniform and healthy (CK-H2O). When the respiratory inhibitor SHAM was added (CK-SHAM), the REC increased, and the RWC of the leaves decreased. However, supplementation with the respiratory enhancer PA (CK-PA) did not cause significant changes in REC and RWC compared to the control. After 9 days of treatment with a 5% CO2 concentration, the leaves exhibited wilting, drooping, yellowing, and even abscission at the apex (T-H2O). The yellowing and abscission of leaves worsened with the addition of SHAM (T-SHAM), while the wilting was alleviated with the addition of PA (T-PA) (Figure 1a). After 6 days of treatment, the REC of the CO2-treated leaves significantly increased by 34.2%, 27.2%, and 52.2% in T-H2O, T-PA, and T-SHAM, respectively (Figure 1b). However, the RWC of T-H2O, T-PA, and T-SHAM significantly decreased by 7.6%, 5.3%, and 8.4%, respectively, after 9 days of CO2 treatment (Figure 1c).
Under normal CO2 conditions, the presence of PA did not have a significant effect on the photosynthetic parameters (Pn, Gs, Ci, and Tr) of the tetraploid R. pseudoacacia’s leaves compared to the control (CK-H2O). However, after supplementation with SHAM on the 6th day, there was a slight fluctuation and decrease in Pn, Gs, and Tr (excluding Ci) compared to the control. When exposed to high CO2, the photosynthetic parameters of the leaves (Pn, Gs, Ci, and Tr) significantly decreased compared to the control (CK-H2O), reaching the lowest values on the 9th day. Pn, Gs, Ci, and Tr were 49.5%, 50.5%, 63%, and 63.7% lower than the control, respectively. After the high CO2 treatment, R. pseudoacacia showed some recovery for three days under normal CO2 conditions, with an increase in the photosynthetic parameters. However, these parameters were consistently lower than the control (Figure 1d–g). In conclusion, the results indicate that the photosynthetic gas parameters of R. pseudoacacia leaves significantly decreased after exposure to high CO2. The addition of the respiration enhancer PA mitigated the inhibition of photosynthesis caused by high CO2, while the addition of the respiration inhibitor SHAM enhanced the inhibition of photosynthesis by high CO2.

2.2. Response of Antioxidant System of Tetraploid R. pseudoacacia to High CO2

To investigate the impact of high CO2 on the antioxidant system of the tetraploid R. pseudoacacia, we initially measured the levels of H2O2 and O2·−. Under normal CO2 conditions, the levels of H2O2 and O2·− decreased after PA treatment compared to CK-H2O, whereas the levels of H2O2 and O2·− increased after SHAM supplementation. After 6 days of high CO2 treatment, the levels of H2O2 and O2·− in the leaves were significantly higher than in CK-H2O. Furthermore, after SHAM supplementation, the levels of H2O2 and O2·− were further increased compared to T-H2O. On the other hand, after PA treatment, the levels of H2O2 and O2·− were slightly decreased compared to T-H2O but were still higher than CK-H2O (Figure 2c,f). In addition, there was no significant difference in the staining intensity of DAB and NBT on the leaves before CO2 treatment. However, after 6 days of high CO2 treatment, the mean staining intensities of DAB and NBT on the leaves were significantly higher compared to CK-H2O. Similarly, the SHAM treatment increased the mean staining intensity, while PA treatment decreased it compared to T-H2O (Figure 2a–f).
SOD and CAT are considered the most significant enzymes in the plant antioxidant system, as their activity levels indicate the extent of damage caused by external factors. Meanwhile, GR and APX play a crucial role in maintaining the balance of the ASA-DHA cycle. APX functions to eliminate the excessive accumulation of H2O2 in a plant, while GR is primarily responsible for converting oxidized glutathione to reduced glutathione. When CO2 conditions are normal, the addition of SHAM slightly enhances the activities of SOD, CAT, GR, and APX. On the other hand, the supplementation of PA does not lead to any significant changes in the activities of these enzymes. However, under high CO2 levels, the activities of SOD, CAT, GR, and APX decrease. Specifically, in T-H2O, T-PA, and T-SHAM, SOD activity decreased by 29.7%, 22.4%, and 54.3%, respectively, CAT activity decreased by 40.3%, 24.1%, and 47.8%, respectively, GR activity decreased by 39.6%, 32.6%, and 46%, respectively, and APX activity decreased by 49.2%, 32.8%, and 62.7%, respectively (Figure 2g–j).

2.3. Changes in Chlorophyll Levels and Fluorescence Parameters

Under normal CO2 conditions, there were no significant differences in the content of Chl a and Chl b, as well as the ratio of Chl a/b, among the CK-H2O, CK-PA, and CK-SHAM groups. However, in the high CO2 treatment group, the levels of Chl a and Chl b increased, but the ratio of Chl a/b decreased compared to CK-H2O after 3 days. In the CK-PA and CK-SHAM groups, there was a significant decrease in the levels of Chl a and Chl b, while the ratio of Chl a/b significantly increased after 6–9 days. Interestingly, the T-SHAM group had lower levels of Chl a and Chl b, whereas the T-PA group had higher levels of Chl a and Chl b (Figure 3a–c).
Under normal CO2 conditions, the leaf chlorophyll fluorescence parameters did not show a significant response to supplementation with SHAM and PA compared to the control. However, after 9 days of high CO2 treatment, Fo and NPQ increased by 15.6% and 63.4%, respectively. Fo and NPQ were higher than in T-H2O after supplementation with SHAM but lower than in T-H2O after supplementation with PA. Upon restoration of normal CO2 conditions, Fo and NPQ returned to the control values (Figure 3d,g). During the high CO2 treatment, Fm, Fv/Fm, qP, and ETR exhibited a decreasing trend and were significantly different compared to CK-H2O. On day 9 of treatment, Fm, Fv/Fm, qP, and ETR decreased by 7.3%, 51.3%, 31% and 56.5%, respectively. Fm, Fv/Fm, qP, and ETR showed a more significant decrease with SHAM supplementation compared to T-H2O, while the addition of the accelerator increased Fm, Fv/Fm, qP, and ETR (Figure 3e,f,h,i). These results indicate that high CO2 has a severe impact on the chlorophyll levels and fluorescence parameters of the tetraploid R. pseudoacacia.

2.4. Effect of High CO2 on Leaf Respiration Parameters

Under normal CO2 conditions, the presence of PA contributed to an increase in Valt and Vt, while the addition of SHAM inhibited the increase in Vcyt, Valt, and Vt. After 6 days of high CO2 treatment, there was a rapid increase in Valt but a significant decrease in Vcyt and Vt compared to CK-H2O. When compared to T-H2O, Vcyt, Valt, and Vt increased with PA supplementation but decreased with SHAM supplementation. Upon restoring normal CO2 conditions, there were no significant differences in Vcyt, Valt, and Vt compared to the control (Figure 4a–c).
The enzyme activities of AOX, complex I, and complex II were increased by PA compared to CK-H2O. No significant changes in the enzyme activities of AOX, complex I, and complex II were observed after supplementation with SHAM under normal CO2 conditions. However, the enzyme activities of AOX, complex I, and complex II were significantly increased under high CO2 treatment. Furthermore, PA supplementation led to a more significant increase in the enzyme activities of AOX, complex I, and complex II compared to T-H2O, while the SHAM treatment also resulted in a significant increase in the enzyme activities of AOX, complex I, and complex II (Figure 4d–f). There were no significant changes in the enzyme activities of complexes III and IV, including after PA and SHAM supplementation, under normal CO2 conditions. However, under high CO2 conditions, the enzyme activities of complexes III and IV were significantly reduced. PA increased the enzyme activities of complexes III and IV compared to T-H2O, while SHAM decreased the enzyme activities of complexes III and IV. After restoration of normal CO2 conditions, the mitochondrial electron transport chain enzyme activities were not significantly different from those of the control (Figure 4g,h).

2.5. Effect of High CO2 on Leaf Stomatal Movement

To investigate the movement of stomata in the tetraploid R. pseudoacacia under high CO2 conditions, we conducted observations on stomatal morphology and counted the stomatal apertures. When exposed to normal CO2 levels, SHAM caused stomata closure, while PA had a lesser effect on the stomata and did not significantly change the stomatal aperture compared to CK-H2O. However, under the high CO2 treatment, T-H2O showed a 67.9% reduction in stomatal apertures compared to CK-H2O. Additionally, supplementation with SHAM further decreased the stomatal apertures by 30% compared to T-H2O, whereas supplementation with PA increased the stomatal apertures by 81.7% (Figure 5).
The fluorescence staining intensity of the guard cells was significantly enhanced after high CO2 treatment, as observed through the use of the fluorescent probes H2DCF-DA and DAF-2DA. This indicates that high CO2 treatment increased the accumulation of ROS and NO in the guard cells. Supplementation with SHAM further increased the fluorescence staining intensity, suggesting an increase in the accumulation of ROS and NO. On the other hand, supplementation with PA decreased the fluorescence staining intensity, indicating a decrease in the accumulation of ROS and NO in the guard cells (Figure 6). In conclusion, the tetraploid R. pseudoacacia regulates stomatal closure in response to a high CO2 environment by inducing the accumulation of ROS and NO in the guard cells.

2.6. Proteomic Analysis of Tetraploid R. pseudoacacia Based on High CO2 Conditions

To investigate the response of the tetraploid R. pseudoacacia to high CO2 levels, we conducted a proteomic analysis on six groups of R. pseudoacacia leaves (CK-H2O, CK-PA, CK-SHAM, T-H2O, T-PA, and T-SHAM). The quality of the samples was confirmed by analyzing the total ion chromatograms, which showed uniform peaks (Figure S1). Additionally, the PCA analysis indicated low variability among samples within each group, as the distribution of the three biological replicates was relatively concentrated. Two-dimensional plots demonstrated significant differences in the principal components between the high CO2 treatment groups and the groups under normal CO2 conditions (Figure S2). A total of 1652 proteins were quantified below the protein threshold of the 1.0% FDR criterion. Furthermore, we randomly selected six DAPs from the proteome sequencing data and examined the variations in their transcript levels using qRT-PCR. The results showed that the abundance variations of the six candidate proteins under high CO2 treatment were consistent with the trend of transcriptional expression (Figure S6).
After conducting a one-way ANOVA test, we screened the differential protein data using Fold Change in Expression (FC) ≥ 1.5-fold as a criterion. To investigate the response mechanism of the tetraploid R. pseudoacacia under high CO2 treatment, we performed a systematic clustering analysis of the differentially abundant proteins (DAPs) (Figure S4). The results revealed a significant change in the expression of DAPs under high CO2 treatment. Under normal CO2 conditions, in the CK-PA/CK-H2O group, we found 50 DAPs, with 11 proteins’ abundances increased and 39 proteins’ abundances decreased (Figure 7b). Similarly, in the CK-SHAM/CK-H2O group, we observed 194 DAPs, with 82 proteins’ abundances increased and 112 proteins’ abundances decreased (Figure 7c). In the T-H2O/CK-H2O group, we identified 504 DAPs, out of which 142 proteins were in elevated abundance, including 3-hydroxybutyryl-CoA dehydrogenase (I3SLI5), glutamine synthetase (I3T8A0), and cytochrome c oxidase subunit Vb (I3SVI7), while 362 proteins were reduced in abundance, including calnexin homolog (I3SD49), glutathione transferase (I3T7D8), and the NAD-dependent epimerase/dehydratase domain-containing protein (I3SMX7) (Figure 7a). In the T-PA/CK-H2O group, we detected 522 DAPs, with 351 proteins, including aquaporin PIP13 (A0A0U2BZJ5), photosystem II protein D1 (A0A1C7D3T6), and class V chitinase CHIT5 (A0A1B1J8Z2), showing a significant reduction in abundance. On the other hand, xylose isomerase (I3S153), NADH-ubiquinone oxidoreductase chain 5 (G9JLS7), cytochrome c oxidase subunit 3 (G9JLU6), and 171 other proteins exhibited a significant increase in abundance (Figure 7d). In the T-SHAM/CK-H2O group, a total of 506 DAPs were identified. Among these, the abundance of 155 proteins, including ATP synthase β-subunit (Q9FUC8), 3-hydroxybutyryl-CoA dehydrogenase (I3SLI5), and 15-cis-phytoene synthase (I3SHI7), was significantly reduced. On the other hand, the relative abundance of 351 proteins, including peroxidase (I3SDX6), ribonuclease T (I3SNB3), and the prolyl 4-hydroxylase alpha subunit domain-containing protein (I3SZ73), was significantly increased (Figure 7e). In the T-PA/T-H2O group, there were a total of 37 DAPs, with 20 proteins’ abundances increased and 17 proteins’ abundances decreased (Figure 7f). Similarly, in the T-SHAM/T-H2O group, there were 35 DAPs, with 16 being upregulated and 19 being downregulated (Figure 7g).
Under normal CO2 conditions, the comparison of CK-PA/CK-H2O and CK-SHAM/CK-H2O revealed 26 overlapping DAPs (Figure S3A). However, under high CO2 treatment, there was a significant increase in the number of DAPs in the tetraploid R. pseudoacacia, with a total of 388 overlapping DAPs in T-PA/CK-H2O and T-SHAM/CK-H2O (Figure S3B). When comparing T-PA/T-H2O and T-SHAM/T-H2O to T-H2O, there were only 19 overlapping DAPs after PA and SHAM treatments at high CO2 concentrations (Figure S3C).

2.7. KOG Functional Annotation and GO Enrichment Analysis of Tetraploid R. pseudoacacia DAPs’ Response to High CO2

The distribution of KOG functions was analyzed, and the results showed that, under high CO2 treatment, the most abundant functions in T-H2O/CK-H2O (52 DAPs) were posttranslational modification, protein turnover, and chaperones; nucleotide transport and metabolism, translation, ribosomal structure and biogenesis, and energy production and conversion included 48, 41, and 35 DAPs, respectively (Figure 8a). In the case of normal CO2 treatment, the DAPs in CK-PA/CK-H2O were mainly associated with general function prediction only, lipid transport and metabolism, posttranslational modification, protein turnover, and chaperones (Figure 8b), whereas the DAPs in CK-SHAM/CK-H2O were mainly distributed in translation, ribosomal structure and biogenesis, posttranslational modification, protein turnover, and energy production and conversion (Figure 8c). Under high CO2 treatment, the DAPs in T-PA/CK-H2O were mainly associated with general function prediction only, posttranslational modification, protein turnover, chaperones, and energy production and conversion (Figure 8d). Similarly, the DAPs in T-SHAM/CK-H2O were mainly related to general function prediction only, posttranslational modification, protein turnover, chaperones, intracellular trafficking, secretion, and vesicular transport (Figure 8e). There was no significant difference in the distribution of KOG function between T-PA/T-H2O and T-SHAM/T-H2O after supplementation with PA and SHAM under high CO2 treatment (Figure 8f,g).
In the GO enrichment analysis, DAPs were categorized into three groups: biological processes, molecular functions, and cellular components. Under normal CO2 conditions, there were fewer DAPs in the CK-PA/CK-H2O group with no significant differential enrichment (Figure S5a). For the CK-SHAM/CK-H2O group, the most enriched biological processes were cellular (128 DAPs) and metabolic (118 DAPs) processes; the most enriched molecular functions were catalytic activity (107 DAPs) and binding (99 DAPs); and the most enriched cellular components were cell parts (142 DAPs) and intracellular parts (135 DAPs) (Figure S5c). Under high CO2 treatment, the most responsive biological processes were cellular (287 DAPs) and metabolic (252 DAPs) processes in the T-H2O/CK-H2O group; the most responsive molecular functions were catalytic activity (259 DAPs) and binding (259 DAPs); and the most responsive cellular components were cell parts (327 DAPs) and intracellular parts (309 DAPs) (Figure S5b). The significantly enriched categories of each GO classification in the T-PA/CK-H2O and T-SHAM/T-H2O groups largely overlapped with those in the T-H2O/CK-H2O group (Figure S5d,e). These results indicate a significant increase in GO-enriched DAPs under high CO2 treatment.

2.8. Pathway Analysis of DAPs in Tetraploid R. pseudoacacia under High CO2 Treatment

To further elucidate the metabolic pathways involved in the DAPs of the tetraploid R. pseudoacacia, DAPs were analyzed for KEGG enrichment. Under normal CO2 conditions, the DAPs in the CK-PA/CK-H2O group were mainly enriched in fatty acid biosynthesis, fatty acid metabolism, and biotin metabolism (Figure 9b). In the CK-SHAM/CK-H2O group, the DAPs were mainly enriched in metabolic pathways, the biosynthesis of secondary metabolites, glutathione metabolism, and carotenoid biosynthesis (Figure 9c). Under high CO2 treatment, the DAPs in the T-H2O/CK-H2O group were significantly enriched in metabolic pathways, carbon metabolism, the biosynthesis of secondary metabolites, the biosynthesis of amino acids, and photosynthesis (Figure 9a). The DAPs in the T-PA/CK-H2O and T-SHAM/CK-H2O groups were significantly enriched in metabolic pathways, the biosynthesis of secondary metabolites, the biosynthesis of amino acids, phenylalanine metabolism, phenylacetone biosynthesis, and glutathione metabolism (Figure 9d,e). Finally, the DAPs in the T-PA/T-H2O and T-SHAM/T-H2O groups were mainly enriched in phenylalanine metabolism, metabolic pathways, the biosynthesis of secondary metabolites, and phenylpropanoid biosynthesis (Figure 9f,g).

3. Discussion

Short-term CO2 enrichment has been found to alleviate the impact of abiotic stresses, such as drought and salt, on plants [47,48]. In the tetraploid R. pseudoacacia, prolonged exposure to high CO2 resulted in a gradual decrease in the relative water content of its leaves over time, ultimately leading to the inhibition of its growth and development. In the present study, we observed the yellowing, wilting, and abscission of leaves, as well as a decrease in the relative water content in the tetraploid R. pseudoacacia at high CO2 concentrations (Figure 1a). This phenomenon could be attributed to a reduction in water flux caused by high CO2 and weakened upward water transport due to a decreased transpiration rate [49,50]. These factors contribute to the recovery of plants once the CO2 concentration is restored (Figure 1c). The magnitude of the REC can indicate the condition of the plant membrane system. When plants face adversity or injuries, the cell membrane may rupture, causing a leakage of cytoplasmic cytosol and consequently increasing the conductivity. The substantial increase in REC observed at high CO2 concentrations suggests that the plant experienced stress injury and cell damage (Figure 1b). PA, known as a facilitator of the alternate respiration pathway, and SHAM, an inhibitor [32,33,35], have shown contrasting effects on the stress symptoms of the tetraploid R. pseudoacacia. PA supplementation alleviated the stress response of the tetraploid R. pseudoacacia to high CO2, whereas SHAM supplementation exacerbated this stress response (Figure 1).
Photosynthesis involves important physiological indicators used to study the impact of abiotic stress on plants. High levels of CO2 can reduce the photosynthetic capacity of plants. This reduction is mainly seen through decreased CO2 assimilation capacity and photochemical activities [10,13,51]. In our study, we observed a decrease in photosynthetic parameters in the tetraploid R. pseudoacacia’s leaves when exposed to high CO2, indicating the inhibition of photosynthetic performance. However, after treatment with PA, the photosynthetic performance showed signs of recovery. After SHAM treatment, photosynthesis in the tetraploid R. pseudoacacia was found to be even more inhibited (Figure 1e–g). This could be due to a decrease in photosynthetic parameters resulting from a reduction in the amount and activity of Rubisco, stomatal conductance, and photorespiration [52,53].
A high concentration of CO2 induces an accumulation of ROS in plants [54,55]. Previous studies have shown that an excessive accumulation of electrons in the PSII leads to an over-reduction in electrons, which in turn causes the electron transport chain to become highly reduced, resulting in the formation of ROS. This oxidative stress not only causes damage to plants but also triggers retrograde signals to the nucleus, leading to responses at the gene expression level [56,57]. In the tetraploid R. pseudoacacia, there was a significant increase in ROS levels compared to the control group under high CO2. This change in ROS levels was observed when the AOX pathway was stimulated or inhibited (Figure 2a–c), which aligns with Dinakar’s study [58]. Algae have developed various antioxidant defense mechanisms to mitigate the oxidative stress caused by H2O2. These mechanisms include enzymatic processes involving the ascorbate–glutathione cycle, glutathione S-transferase (GST), glutathione peroxidase (GPX), and CAT. Additionally, nonenzymatic mechanisms related to carotenoids and glutathione also play a role [59,60,61]. Most ROS-scavenging enzymes, such as CAT, SOD, and APX, are predominantly bound to cystoid membranes and exhibit similarities. When plants experience stress, they promptly remove ROS to prevent toxicity caused by ROS diffusion into the matrix [61]. To maintain ROS homeostasis, the activity of antioxidant enzymes is typically enhanced to eliminate excess ROS, thus safeguarding the plant from oxidative damage and enhancing its stress tolerance [62,63]. However, this mechanism is effective against mild oxidative damage, and if the damage surpasses the plant’s tolerance threshold, the inability to eliminate ROS will hinder the functioning of antioxidant enzymes [64,65]. In this study, high concentrations of CO2 led to a significant increase in ROS levels in the tetraploid R. pseudoacacia, exceeding the capacity of the antioxidant system and resulting in the suppression of SOD, CAT, GR, and APX activities, ultimately causing damage to the plant (Figure 2d–f).
Chlorophyll content can have a direct impact on a plant’s capacity to capture light energy. The decrease in chlorophyll content can be attributed to two factors: the blockage of chlorophyll synthesis and the degradation of chlorophyll [66]. In our study, we observed that the levels of both Chl a and b initially increased and then decreased under high CO2 treatments (Figure 3a,b). This finding aligns with previous research and suggests that high concentrations of CO2 may inhibit chlorophyll synthesis [67]. We investigated the impact of an elevated CO2 concentration on the activity of PSII using the chlorophyll fluorescence technique. The intensity of chlorophyll fluorescence is an indicator of the redox state of the main receptor Q in PSII [68]. Our findings revealed that the treatment with high CO2 levels resulted in an increase in Fo but a significant decrease in Fm and Fv/Fm in the tetraploid R. pseudoacacia (Figure 3d–f). This suggests that the activity of PSII was hindered, leading to an excessive accumulation of electrons in PSII. Additionally, qP and NPQ represent the capacities for photochemical and nonphotochemical excitation quenching, respectively, and the quantum yields of these two types of fluorescence quenching are interconnected [69]. A decrease in ETR signifies a decline in photosynthetic activity. In the tetraploid R. pseudoacacia, photosynthesis is hindered, disrupting the balance of the system. When the plant is under stress, PSII is inhibited (Figure 3), resulting in the generation of a chlorophyll triplet state and singlet oxygen (1O2), which ultimately leads to the accumulation of ROS.
The regulation of the mitochondrial electron transport chain, which includes the cytochrome respiratory (COX) and AOX respiratory pathways, is influenced by abiotic stress. Under high CO2 conditions, AOX respiration is upregulated in plant mitochondria to support normal plant physiological activities [70]. Research has demonstrated the significance of the AOX pathway in A. thaliana in response to elevated CO2 [71]. It was observed that the rate of cytochrome respiration in the tetraploid R. pseudoacacia decreased, while the rate of AOX respiration increased under high CO2 stress. However, overall respiration was inhibited. AOX respiration, which is a SHAM-sensitive type of respiration, was stimulated by PA treatment and had a significant effect on AOX [32,33,35]. PA treatment notably enhanced the AOX respiration pathway, while the response of AOX respiration in the tetraploid R. pseudoacacia to high CO2 was inhibited by SHAM (Figure 4d). In contrast, the mitochondrial transfer chain (ETC) complexes (I, II, III, and IV) exhibited distinct responses under high CO2. The activities of complexes I and II were decreased, while the activities of complexes III and IV were downregulated (Figure 4e–h). This alteration in activity levels could be attributed to mitochondrial damage resulting from the accumulation of ROS.
Low CO2 promotes stomatal opening, while high CO2 induces stomatal closure. Stomata, which serve as the primary site for gas exchange, can adjust the pore size in response to changes in the external environment. Studies have shown that high CO2 stimulates an increase in the concentration of free Ca2+ in guard cells, leading to the induction of stomatal closure [14,15,16,17]. Transient changes in the pH and membrane potential of guard cells occur alongside stomatal closure when the CO2 concentration is increased [72]. Stomata control the size of pores in the leaf in response to CO2 concentration, and elevated CO2 also inhibits stomatal development [73,74]. Previous studies have demonstrated that high levels of CO2 can induce stomatal closure, with ROS playing a crucial role in this regulatory process [22,75]. NO plays a role in both ABA- and Ca2+-mediated stomatal closure, which is an important participant in the stomatal closure process [24,25,76]. When compared with normal CO2, high CO2 significantly induces stomatal closure. PA alleviates this process, while SHAM exacerbates the degree of closure. Additionally, SHAM also causes a slight closure of stomata at normal CO2 concentrations. Through fluorescence staining, it was observed that CO2-induced stomatal closure is accompanied by a large accumulation of H2O2 and NO in the guard cells (Figure 5 and Figure 6). This provides evidence for the crucial role of ROS and NO in the stomatal closure of the tetraploid R. pseudoacacia in response to high CO2 stress.
Quantitative proteomics analysis is a crucial technology in proteomics research [77]. We employed DIA quantitative proteomic analysis to investigate protein alterations in the tetraploid R. pseudoacacia under high CO2 treatment. The identified proteins were found to be involved in significant pathways, including metabolic pathways, carbon metabolism, the biosynthesis of secondary metabolites, the biosynthesis of amino acids, and photosynthesis (Figure 7, Figure 8 and Figure 9). Additionally, we screened several key proteins in all DAPs in response to high CO2, including ribulose diphosphate carboxylase-related subunits (A0A0F6Y5S8, C0J370), photosystems I-II-related proteins (A0A898CWR3, A0A898CTK1, A0A898CW46, A0A898CTN3, A0A898CW14), antioxidant-related proteins (I1KP94, A0A0R4J3P1, K7KDM7, A0A0R0ERX9), cytochrome-related proteins (A0A898CTT0, I1MWP7, A0A0R0H975), and NADPH-related proteins (A0A0R4J3V9, I1LT63) (Table S2).
In C3 plants, glucose is primarily biosynthesized through the Calvin cycle. A crucial step in this process is the formation of glycerate-3-phosphate (3PGA) through the action of 1,5-bisphosphate ribulose carboxylase using CO2 [28]. Rubiscos, which are photosynthetic CO2-fixing enzymes, are activated by Rubisco activase (Rcas). However, if the N-terminus of the large subunit is missing, Rubiscos are not activated [78]. The involvement of carbon dioxide is also essential for the activation of Rubiscos [79]. In this study, we observed a significant downregulation of many 1,5-bisphosphate ribulose carboxylase-related subunits (I6QMJ2, A0A0F6Y5S8, C0J370, A0A898CTM2) under high CO2 treatment. This downregulation resulted in the inability of Rubiscos to be activated under light and CO2 conditions, leading to a decrease in the photosynthesis rate. Furthermore, the inhibition of photosynthesis was also attributed to the impact on photosystem protein synthesis, electron transfer between photosystems, and chlorophyll binding [80,81,82]. Pigment–protein complexes CP43 and CP47 are known to transfer excitation energy from the outer antenna of photosystem II to the photochemical reaction center, as well as being involved in the process of water oxidation [83,84]. The excess energy absorbed by the photosystem I complex is transferred to P700 via chlorophyll, thus protecting the pigment–protein complex from photodestruction [85]. The photosystem II D1/D2 complex proteins catalyze the oxidation of water due to a high REDOX potential [86]. Light trapping and the conversion of light energy to chemical energy occur through the iron–sulfur center in photosystem I [87,88]. In our study, we observed a significant downregulation of various photosystem constitutive proteins in the tetraploid R. pseudoacacia. These proteins include photosystem II CP43/47 reaction center proteins (A0A6H0EHP1, A0A1C7D4D5), photosystem I P700 chlorophyll a apoprotein A1/A2 (A0A1C7D3U7, A0A1C7D4C6, A0A6H0EHR0), photosystem I reaction center subunit II (I3STB2), photosystem II D1/D2 proteins (A0A1C7D3T6, A0A1C7D3V4), electron-transfer-mediating protein cytochrome f (A0A1C7D3W5), plastoquinone–plastocyanin reductase (I3RZ73, I1LUB3), the photosystem I iron–sulfur center (A0A1C7D3Z6), and the chlorophyll a–b binding protein (I3SIW2). This downregulation resulted in the inhibition of electron transfer and light capture, ultimately leading to the blockage of photosynthesis in the tetraploid R. pseudoacacia.
Glutathione (GSH) plays a crucial role in various biological processes. When GSH is oxidized to oxidized glutathione (GSSG) by GPX, it catalyzes the reduction of H2O2. Additionally, GR helps in regenerating GSH. Additionally, GSH has been found to effectively increase the antioxidant capacity of plants and enhance the activity of respiratory enzymes [89]. In certain studies, advanced GSH treatment has proven to be effective in preventing broccoli fermentation in a high CO2 atmosphere. This is achieved by promoting the AsA–GSH cycle and the electron (ETC) pathway [78]. Exogenous PA can maintain intracellular glutathione levels in H2O2-treated cells and enhance their antioxidant capacity [90]. The presence of GSH is vital for copepods in defending against seawater acidification induced by varying CO2 concentrations [91]. In this study, the expression of GPX (I3SK85, I1KP94, I1MX60, C6SY48) was significantly decreased by high CO2 treatment. However, this decrease was reversed when PA was applied, indicating that the antioxidant capacity of the tetraploid R. pseudoacacia was restored. The decrease in GPX expression resulted in the accumulation of ROS. GR plays a crucial role in the interconversion between oxidized and reduced glutathione, facilitating the recirculation of GSH. In soybeans, enhancing GR can effectively improve the removal of ROS [92]. In the context of this experiment, high CO2 levels led to the downregulation of GR and diminished the regeneration ability of GSH. However, the application of PA facilitated the conversion between GSH and GSSG under high CO2 conditions, which enhanced the scavenging of H2O2 and alleviated oxidative damage.

4. Materials and Methods

4.1. Plant Materials and Stress Treatment

The plant materials comprised the tetraploid R. pseudoacacia, which was two years old. In the spring, bare-root seedlings were planted in uniform-sized pots (diameter × height × bottom diameter = 20 × 28 × 20 cm), and seedlings with uniform growth were chosen for CO2 treatment after 2 months of growth under normal conditions. The potted seedlings were placed in a light incubator for 3 days for pre-cultivation to acclimate to the new environment. The soil moisture was maintained at 70% throughout, and the culture conditions were as follows: 16 h of light, 8 h of darkness, temperature set at 25 °C, air humidity at 70%, and light intensity at 6000 Lux. The seedlings undergoing the high CO2 treatment were divided into a control group (natural conditions, a CO2 concentration of about 0.031%) and treatment groups (a CO2 concentration of 5%), and the seedlings were subjected to high CO2 treatment for 9 days before being incubated again under normal CO2 concentration for 3 days. On day 6 of CO2 treatment, a respiration accelerator (pyruvic acid, 0.1 mM) and a respiration inhibitor (salicylhydroxamic acid, 1 mM) were sprayed on the leaves of the tetraploid R. pseudoacacia every 3 h. Morphological observations on leaves were made after 0 d, 3 d, 6 d, 9 d, and 12 d. For each treatment, leaves were collected for physiological and proteomic measurements, with three or six biological replicates.
Here, CK-H2O: untreated control group; CK-PA: pyruvic acid-treated group; CK-SHAM: salicylhydroxamic acid-treated group; T-H2O: CO2-treated group; T-PA: CO2 co-treated with PA; and T-SHAM: CO2 co-treated with SHAM.

4.2. Determination of Relative Water Content and Relative Electrical Conductivity

After 0.5 g of fresh leaves (FWs) moistened in distilled water for 4 h, the surface water was removed, weighed, and recorded as TW, and then the leaves were dried in a drying oven and recorded as DW. Relative water content: RWC (%) = (FW − DW)/(TW − DW) × 100.
A total of 0.1 g of evenly chopped leaves were poured in 5 mL of ddH2O and shaken at 160 rpm for 1 h. The electrical conductivity was measured and recorded as L2 using a DDS-IIA conductivity meter (INESA Scientific Instruments Co., Shanghai, China), and the control of ddH2O was recorded as L1. After cooling to ambient temperature, the samples were immersed in a boiling water bath for 15 min, and the electrical conductivity was measured and recorded as L3. Relative electrical conductivity (%) = (L2 − L1)/(L3 − L1) × 100.

4.3. Determination of Chlorophyll Content and Photosynthetic and Fluorescence Parameters

The chlorophyll content of the fourth and fifth leaves from the end, located in the middle of the acacia branches, was determined. A total of 0.2 g of leaves were immersed in the extraction solution (80% acetone/ethanol/ddH2O = 4.5:4.5:1). The leaves were then kept in the dark for 24 h. The optical density at wavelengths of 470 nm, 645 nm, and 663 nm was measured. Chlorophyll a (Chl a) = 12.72A663 − 2.59A645; chlorophyll b (Chl b) = 22.88A645 − 4.67A663; and Chl a/b = chlorophyll a/b.
The photosynthetic parameters of leaves were determined using a Li-COR 6400 photosynthesis system (LI-COR, Lincoln, NE, USA). Six leaves per group were analyzed, and the measurements were repeated three times. These parameters included the net photosynthetic rate (Pn), the transpiration rate (Tr), the intercellular carbon dioxide concentration (Ci), and the stomatal conductance (Gs). The measurements were conducted under specific conditions, including a temperature of 24 °C, a photon flux density (PFD) of 90 µmol m−2·s−1, a relative humidity of 70%, and an ambient CO2 concentration of 400 µmol CO2 mol−1.
The chlorophyll fluorescence parameters of leaves were measured using an FC800-O fluorescence imaging system (FluorCam, Drásov, Czech Republic) [93]. The leaves were dark-adapted for 30 min, following the previously described method. The data collected included initial fluorescence yield (Fo), maximum fluorescence yield (Fm), photochemical quenching (qP), the maximum photochemical efficiency of photosystem II (PSII) (Fv/Fm), nonphotochemical quenching (NPQ), and the relative electron transfer rate (ETR). A total of six leaves were measured for each treatment group.

4.4. Determination of Leaf Respiration Parameters

The leaf respiration rate was determined using the OXYTHE-RM oxygen electrode. The total respiration rate (Vt) was determined by incubating 0.1 g leaves of R. pseudoacacia from the same location in a reaction medium containing 2 mM HEPES, 10 mM MES (pH 10.7), and 2 mM CaCl2 for 30 min in the dark. To determine the residual respiration (Vres), SHAM (200 mM) and KCN (200 mM) were added to the assay. Vres was then subtracted from the respiratory activity observed when SHAM was added alone to obtain the activity of the major cytochrome pathway (Vcyt). Similarly, Vt was subtracted from the respiratory activity observed when SHAM was added alone to obtain the actual activity of the alternative pathway (Valt).
The activity of mitochondrial electron transport chain complex I-IV (complex I-IV) and AOX was measured in the central leaves of R. pseudoacacia using the Plant Mitochondrial Respiratory Chain Complex I-IV and Plant AOX Activity Kit (Hengyuan Biologicals, Shanghai, China).

4.5. Stomatal Movement

Under controlled light and temperature conditions, fresh leaves were collected from plants grown under normal conditions and with high CO2 treatments. The upper epidermis and leaf pulp of both the proximal and distal axes of the leaf blades were gently scraped. Stomatal morphology was then observed randomly using a microscope. More than 100 stomata were selected for each treatment sample. The length and width of the stomatal images were measured using ImageJ v1.8.0 to calculate the stomatal openness.

4.6. Fluorescent Probe Staining

Fresh leaves were collected from plants under both normal conditions and high CO2 treatments after 0, 6, and 12 days. Leaf blades, excluding the main veins, were cut into 1 cm pieces. These pieces were then immersed in 50 mM H2DCF-DA or 0.1 mM DAF-2DA for 30 min in the absence of light. Subsequently, they were photographed using a fluorescence microscope in 10 randomly selected fields of view. The average fluorescence intensity was calculated using ImageJ.

4.7. Histochemical Staining

The leaves were fully immersed in a staining solution containing 10 mg·mL−1 DAB and NBT. To eliminate air bubbles, a vacuum pump was used for 30 min. Subsequently, the leaves were stained under dark conditions for 24 h. Chlorophyll was completely removed by submerging the leaves in a decolorizing solution. The leaves were then photographed and recorded, and the average fluorescence intensity was measured using ImageJ.

4.8. Determination of H2O2 and O2·− Content

A total of 0.5 g of leaves were homogenized in 0.1% TCA in an ice bath. The homogenate was then centrifuged at 12,000 rpm for 10 min. The extract was mixed with phosphate buffer solution (PBS) (pH 7.5) and KI (1 M). The absorbance of the mixture was measured at 390 nm.
Similarly, a total of 0.5 g of leaves were homogenized in hydroxylamine hydrochloride in an ice bath and then centrifuged at 12,000 rpm for 10 min. The supernatant obtained was incubated with 1-Naphthylamine and aminobenzenesulfonic acid for 2 min at room temperature. Subsequently, the absorbance of the solution was measured at 520 nm [94].

4.9. Determination of Antioxidant Enzyme Activity

Crude enzyme extract was prepared by grinding 0.5 g of leaves in liquid nitrogen and centrifuging in PBS at 12,000 rpm for 10 min. A mixture of 20 μL of extract, 50 mM PBS (pH 7.8), 100 μM ethylene diamine tetraacetic acid (EDTA), and 10 mM pyrogallic acid was thoroughly mixed. Superoxide dismutase (SOD) activity was measured by UV spectrophotometry at 420 nm. Another mixture of 0.2 mL of extract, 50 mM PBS, and ddH2O was thoroughly mixed, and the reaction was carried out in a water bath at 25 °C for 10 min; catalase (CAT) activity was measured at 240 nm. Additionally, 0.05 mL of extract, PBS, and 5 mM ASA were mixed thoroughly. This mixture was then followed by the addition of 0.1 mM H2O2, and the activity of ascorbate peroxidase (APX) was measured at 290 nm. Furthermore, a mixture of 10 μL of extract, 100 mM PBS, 2 mM EDTA, and 0.5 mM glutathione was mixed thoroughly. This mixture was then followed by the addition of 0.2 mM NADPH, and the activity of glutathione reductase (GR) was detected at 340 nm [45].

4.10. Quantitative Proteomics Analysis

Six sets of samples were pretreated using the iST Sample Pretreatment Kit (PreOmics, Planegg, Germany) and stored at −80 °C under vacuum conditions. The samples were crushed in liquid nitrogen and then mixed with buffer (1:10) and protease inhibitors. After vortexing for 10 min, the samples were vortexed for another 10 min with an equal volume of Tris-saturated phenol (pH 8.0). Subsequently, they were centrifuged at 12,000 rpm for 20 min at 4 °C, and the phenol phase was combined with buffer and vortexed before centrifugation. The precipitate was then treated with pre-cooled ammonium acetate–methanol solution, precipitated overnight at −20 °C, and finally centrifuged at 12,000 rpm for 20 min at 4 °C. The supernatant was discarded, and the precipitate was washed twice with 90% acetone. The precipitate was then suspended in an appropriate volume of lysate to solubilize the sample proteins, followed by centrifugation at 12,000 rpm for 20 min at 4 °C to collect the supernatant. This step was repeated to ensure maximum collection of the supernatant. Each sample was then suspended in 30 μL of solvent A, and 1 μL of 10×iTR peptide was added to 9 μL of each sample. After thorough mixing, the samples were subjected to nanoliquid chromatography. For tandem mass spectrometry analysis, 4 μL of the samples were taken. The separation of the sample occurred over a 90 min gradient with a column flow rate of 600 nL·min−1 and a column temperature of 55 °C. The gradient started at 4% B-phase and was equilibrated for 4 min, followed by a nonlinear gradient increase to 30% over 80 min. Subsequently, the gradient increased to 90% in 2 min and was held for 8 min.
The mass spectrometry conditions were as follows: (1) MS: scan range (m/z): 350–1500; resolution: 120,000; normalized AGC target: 100%; and maximum injection time: 50 ms; (2) HCD-MS/MS: resolution: 30,000; normalized AGC target: 600%; maximum injection time: 90 ms; and collision energy: 35; (3) variable window acquisition was used with 60 windows set up, and overlapping serial ports were set up with 1 m·z−1 overlap per window.
Data analysis: DIA data were analyzed using Spectronaut 15.0 default parameters (Omicsolution Co., Ltd., Shanghai, China). The sequence database used was the uniprot-robinioid clade database, and trypsin enzymatic digestion was applied. The search library parameters included a fixed modification, carbamidomethyl (C), and a variable modification, methionine oxidation. The criteria for protein characterization were a precursor threshold of 1.0% false discovery rate (FDR) and a protein threshold of a 1.0% FDR. To generate the decoy database, a mutation strategy was employed, similar to perturbing a random number of amino acid sequences. Spectronaut 15.0 performed auto-correction, and a local normalization strategy was implemented for data normalization. Protein group quantification was conducted by calculating the average of the peak areas of the first 3 peptides below 1.0% FDR. For differential screening, proteins with a p adj value < 0.05 and |fold change| > 1.5 were considered significant after analysis by a one-way ANOVA test.

4.11. Functional Annotation and Cluster Analysis

An unsupervised principal component analysis (PCA) was conducted on the entire dataset to provide a comprehensive overview of the quantitative proteomics data, assessing its relevance and reproducibility. Volcano plots were then analyzed using a one-way ANOVA test, with a significance threshold of a p adj value < 0.05. To identify differentially abundant proteins (DAPs), we considered a fold change greater than 1.5. For the functional annotation and enrichment analysis of the DAPs, we utilized the GO (Gene Ontology), KOG (Eukaryotic Orthologous Groups), and KEGG (Kyoto Encyclopedia of Genes and Genomes) databases, which helped us understand the functions, distributions, and pathways associated with these DAPs.

4.12. Data Submission

MS data were translated to PRIDE XML using the PRIDE Submission Tool Version 2.7.3. A total of 34 PRIDE XML files were submitted to the ProteomeXchange repository following the ProteomeXchange submission guidelines [95]. The data were deposited under the identifier PXD047363.

4.13. Quantitative Real-Time Polymerase Chain Reaction (qRT-PCR)

Samples were ground to powder in liquid nitrogen, and total RNA was isolated using the Omega Plant RNA Kit (Shanghai Yuanmu Biotechnology Co., Shanghai, China) following the manufacturer’s instructions. cDNA was prepared with the PrimeScript Reverse Transcriptase Kit (Takara, Shiga, Japan). qPCR was performed in a LightCycler 480 System (Roche, Indianapolis, IN, USA) with LightCycler 480 SYBR Green I Master Mix and specific primers (Supplementary Table S1). Three biological replicates were used for each sample. The 18SRNA gene was used as an internal control. Relative gene expression levels were calculated using the ΔΔCt method after normalization to the reference gene 18SRNA.

4.14. Statistical Analysis

All experiments were performed with at least three to six biological replicates; experimental data were analyzed with SPSS 25.0; and to determine statistical significance, asterisks indicated significant differences (Student’s t-test; * p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).

5. Conclusions

In conclusion, this study investigated the response mechanism of the tetraploid R. pseudoacacia to high CO2 concentrations through physiology and proteomics. The findings indicate that high CO2 had a negative impact on the normal growth of the tetraploid R. pseudoacacia. Specifically, the photosynthesis and respiration of the tetraploid R. pseudoacacia’s leaves decreased under the high CO2 treatment compared to normal conditions, leading to an excessive accumulation of ROS. Additionally, the activities of the antioxidant enzymes SOD, CAT, GR, and APX were reduced, and the levels of chlorophyll and chlorophyll fluorescence parameters were significantly diminished. Furthermore, the accumulation of ROS and NO in guard cells induced stomatal closure. The proteomic analysis revealed that the DAPs were primarily enriched in metabolic pathways, the biosynthesis of secondary metabolism, and amino acid biosynthesis pathways under high CO2 treatment. Notably, the key DAPs were primarily associated with ribulose diphosphate carboxylase-related subunits, photosystem I-II-related proteins, antioxidant-related proteins, and cytochrome-related proteins. Overall, this study provides valuable insights into the adaptation of the tetraploid R. pseudoacacia to high CO2 environments. However, further research is required to validate the target proteins.

Supplementary Materials

The supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ijms25105262/s1.

Author Contributions

Conceptualization, J.L. and F.M.; methodology, J.L.; validation, J.L. and S.Z.; formal analysis, P.L. and L.G.; data curation, S.Z.; writing—original draft preparation, J.L. and S.Z.; writing—review and editing, F.M.; funding acquisition, X.Z. and F.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant numbers 32071728 and 31170568.

Data Availability Statement

MS data were translated to PRIDE XML using the PRIDE Submission Tool Version 2.7.3. A total of 34 PRIDE XML files were submitted to the ProteomeXchange repository following the ProteomeXchange submission guidelines. The data were deposited under the identifier PXD047363.

Acknowledgments

The authors would like to acknowledge the technical support from the Analysis and Testing Center of Northeast Forestry University.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, H.; Zhang, S.K.; Song, W.Q.; Tigabu, M.; Fu, M.; Xue, H.F.; Sun, A.R.; Zhao, M.H.; Cai, K.W.; Li, Y.; et al. Climate response of radial growth and early selection of Larix olgensis at four trials in northeast China. Dendrochronologia 2022, 73, 125955. [Google Scholar] [CrossRef]
  2. Cubasch, U.; Wuebbles, D.; Chen, D.; Facchini, M.C.; Frame, D.; Mahowald, N.; Winther, J.G. Introduction. In Climate Change 2013: The Physical Science Basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2013; pp. 95–123. [Google Scholar]
  3. Morin, F.; André, M.; Betsche, T. Growth Kinetics, Carbohydrate, and Leaf Phosphate Content of Clover (Trifolium subterraneum L.) after Transfer to a High CO2 Atmosphere or to High Light and Ambient Air 1. Plant Physiol. 1992, 99, 89–95. [Google Scholar] [CrossRef]
  4. Siriphanich, J.; Kader, A.A. Changes in Cytoplasmic and Vacuolar pH in Harvested Lettuce Tissue as Influenced by CO2. J. Am. Soc. Hortic. Sci. 1986, 111, 73–77. [Google Scholar] [CrossRef]
  5. Larrigaudiere, C.; Pintoó, E.; Lentheric, S.; Vendrell, M. Involvement of oxidative processes in the development of core browning in controlled-atmosphere stored pears. J. Hortic. Sci. Biotechnol. 2001, 76, 157–162. [Google Scholar] [CrossRef]
  6. Shuyun, Y.; Qingguo, Z.; Yuelin, J.; Fengwen, W.; Xiaofei, Z. Effect of high CO2 density on photosynthesis speed of Pinus massoniana. J. Anhui Agric. Univ. 2006, 33, 100–104. [Google Scholar]
  7. Jacotot, A.; Marchand, C.; Gensous, S.; Allenbach, M. Effects of elevated atmospheric CO2 and increased tidal flooding on leaf gas-exchange parameters of two common mangrove species: Avicennia marina and Rhizophora stylosa. Photosynth. Res. 2018, 138, 249–260. [Google Scholar] [CrossRef] [PubMed]
  8. Rao, M.V.; Hale, B.A.; Ormrod, D.P. Amelioration of Ozone-Induced Oxidative Damage in Wheat Plants Grown under High Carbon Dioxide (Role of Antioxidant Enzymes). Plant Physiol. 1995, 109, 421–432. [Google Scholar] [CrossRef]
  9. Maroco, J.P.; Edwards, G.E.; Ku, M.S. Photosynthetic acclimation of maize to growth under elevated levels of carbon dioxide. Planta 1999, 210, 115–125. [Google Scholar] [CrossRef] [PubMed]
  10. Tomimatsu, H.; Sakata, T.; Fukayama, H.; Tang, Y. Short-term effects of high CO2 accelerate photosynthetic induction in Populus koreana × trichocarpa with always-open stomata regardless of phenotypic changes in high CO2 growth conditions. Tree Physiol. 2019, 39, 474–483. [Google Scholar] [CrossRef]
  11. Zhou, Y.; Han, S.; Liu, Y.; Jia, X. Stomatal response of Pinus sylvestriformis to elevated CO2 concentrations during the four years of exposure. J. For. Res. 2005, 16, 15–18. [Google Scholar]
  12. Chen, J.; Hu, Y.; Yan, R.; Hu, H. Effect of high carbon dioxide injury on the physiological characteristics of fresh-cut garlic scapes. Sci. Hortic. 2019, 250, 359–365. [Google Scholar] [CrossRef]
  13. Niu, Q.; Huang, D.; Aierken, Y.; Chen, C. Effects of CO2 and Temperature on Canopy Photosynthesis of Muskmelon. Acta Hortic. Sin. 2006, 33, 272–277. [Google Scholar]
  14. Schulze, S.; Dubeaux, G.; Ceciliato, P.H.O.; Munemasa, S.; Nuhkat, M.; Yarmolinsky, D.; Aguilar, J.; Diaz, R.; Azoulay-Shemer, T.; Steinhorst, L.; et al. A role for calcium-dependent protein kinases in differential CO2- and ABA-controlled stomatal closing and low CO2-induced stomatal opening in Arabidopsis. New Phytol. 2020, 229, 2765–2779. [Google Scholar] [CrossRef] [PubMed]
  15. Schwartz, A. Role of Ca2+ and EGTA on Stomatal Movements in Commelina communis L. Plant Physiol. 1985, 79, 1003–1005. [Google Scholar] [CrossRef] [PubMed]
  16. Schwartz, A.; Ilan, N.; Grantz, D.A. Calcium Effects on Stomatal Movement in Commelina communis L. Plant Physiol. 1988, 87, 583–587. [Google Scholar] [CrossRef] [PubMed]
  17. Webb, A.A.R.; Mcainsh, M.R.; Mansfield, T.A.; Hetherington, A.M. Carbon dioxide induces increases in guard cell cytosolic free calcium. Plant J. 2010, 9, 297–304. [Google Scholar] [CrossRef]
  18. Haworth, M.; Killi, D.; Materassi, A.; Raschi, A. Coordination of stomatal physiological behavior and morphology with carbon dioxide determines stomatal control. Am. J. Bot. 2015, 102, 677–688. [Google Scholar] [CrossRef] [PubMed]
  19. Ries, A.B. Carbonic anhydrases function as mediators of CO-induced stomatal movements and regulators of stomatal development in Arabidopsis thaliana. Diss. Theses-Gradworks 2009, 59, 253–257. [Google Scholar]
  20. He, J.; Zhang, R.X.; Peng, K.; Tagliavia, C.; Li, S.; Xue, S.; Liu, A.; Hu, H.; Zhang, J.; Hubbard, K.E.; et al. The BIG protein distinguishes the process of CO2-induced stomatal closure from the inhibition of stomatal opening by CO2. New Phytol. 2018, 218, 232–241. [Google Scholar] [CrossRef]
  21. Muhlemann, J.K.; Younts, T.L.B.; Muday, G.K. Flavonols control pollen tube growth and integrity by regulating ROS homeostasis during high-temperature stress. Proc. Natl. Acad. Sci. USA 2018, 115, E11188–E11197. [Google Scholar] [CrossRef]
  22. Sierla, M.; Waszczak, C.; Vahisalu, T.; Kangasjärvi, J. Reactive Oxygen Species in the Regulation of Stomatal Movements. Plant Physiol. 2016, 171, 1569–1580. [Google Scholar] [CrossRef] [PubMed]
  23. He, J.J.; Zhang, R.X.; Kim, D.S.; Sun, P.; Liu, H.G.; Liu, Z.M.; Hetherington, A.M.; Liang, Y.K. ROS of Distinct Sources and Salicylic Acid Separate Elevated CO2-Mediated Stomatal Movements in Arabidopsis. Front. Plant Sci. 2020, 11, 542. [Google Scholar] [CrossRef] [PubMed]
  24. Yang, J.; Matsumoto, Y.; Etoh, T.; Iwai, S. Nitric oxide (NO)-dependent and NO-independent signaling pathways act in ABA-inhibition of stomatal opening. Plant Signal. Behav. 2008, 3, 131–132. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, X.; Zhang, S.; Lou, C. Involvement of Ca2+ in Stomatal Movements of Vicia faba L. Regulated by Nitric Oxide. J. Plant Physiol. Mol. Biol. 2003, 29, 342–346. [Google Scholar]
  26. Song, Y.; Miao, Y.; Song, C.-P. Behind the scenes: The roles of reactive oxygen species in guard cells. New Phytol. 2014, 201, 1121–1140. [Google Scholar] [CrossRef] [PubMed]
  27. Singh, R.; Parihar, P.; Singh, S.; Mishra, R.K.; Singh, V.P.; Prasad, S.M. Reactive oxygen species signaling and stomatal movement: Current updates and future perspectives. Redox Biol. 2017, 11, 213–218. [Google Scholar] [CrossRef] [PubMed]
  28. Waheeda, K.; Kitchel, H.; Wang, Q.; Chiu, P.L. Molecular mechanism of Rubisco activase: Dynamic assembly and Rubisco remodeling. Front. Mol. Biosci. 2023, 10, 1125922. [Google Scholar] [CrossRef]
  29. Loughran, T.F.; Boysen, L.; Bastos, A.; Hartung, K.; Havermann, F.; Li, H.; Nabel, J.; Obermeier, W.A.; Pongratz, J. Past and Future Climate Variability Uncertainties in the Global Carbon Budget Using the MPI Grand Ensemble. Glob. Biogeochem. Cycles 2021, 35, e2021GB007019. [Google Scholar] [CrossRef]
  30. Ikkonen, E.N.; Shibaeva, T.G.; Sherudilo, E.G.; Titov, A.F. Effect of Continuous Lighting on Mitochondrial Respiration in Solanacea Plants. Russ. J. Plant Physiol. 2022, 69, 114. [Google Scholar] [CrossRef]
  31. McDonald, A.E. Unique opportunities for future research on the alternative oxidase of plants. Plant Physiol. 2023, 191, 2084–2092. [Google Scholar] [CrossRef]
  32. Torrentino-Madamet, M.; Almeras, L.; Travaille, C.; Sinou, V.; Pophillat, M.; Belghazi, M.; Fourquet, P.; Jammes, Y.; Parzy, D. Proteomic analysis revealed alterations of the Plasmodium falciparum metabolism following salicylhydroxamic acid exposure. Res. Rep. Trop. Med. 2011, 2, 109–119. [Google Scholar] [CrossRef] [PubMed]
  33. Mallo, N.; Lamas, J.; Leiro, J.M. Evidence of an Alternative Oxidase Pathway for Mitochondrial Respiration in the Scuticociliate Philasterides dicentrarchi. Protist 2013, 164, 824–836. [Google Scholar] [CrossRef] [PubMed]
  34. Kirimura, K.; Yoda, M.; Shimizu, H.; Sugano, S.; Mizuno, M.; Kino, K.; Usami, S. Contribution of cyanide-insensitive respiratory pathway, catalyzed by the alternative oxidase, to citric acid production in Aspergillus niger. Biosci. Biotechnol. Biochem. 2000, 64, 2034–2039. [Google Scholar] [CrossRef]
  35. Onda, Y.; Kato, Y.; Abe, Y.; Ito, T.; Ito-Inaba, Y.; Morohashi, M.; Ito, Y.; Ichikawa, M.; Matsukawa, K.; Otsuka, M.; et al. Pyruvate-sensitive AOX exists as a non-covalently associated dimer in the homeothermic spadix of the skunk cabbage, Symplocarpus renifolius. Febs Lett. 2007, 581, 5852–5858. [Google Scholar] [CrossRef]
  36. Umbach, A.L.; Wiskich, J.T.; Siedow, J.N. Regulation of alternative oxidase kinetics by pyruvate and intermolecular disulfide bond redox status in soybean seedling mitochondria. FEBS Lett. 1994, 348, 181–184. [Google Scholar] [CrossRef] [PubMed]
  37. Carré, J.E.; Affourtit, C.; Moore, A.L. Interaction of purified alternative oxidase from thermogenic Arum maculatum with pyruvate. Febs Lett. 2011, 585, 397–401. [Google Scholar] [CrossRef] [PubMed]
  38. Sun, Y.R.; Ma, W.T.; Xu, Y.N.; Wang, X.M.; Li, L.; Tcherkez, G.; Gong, X.Y. Short- and long-term responses of leaf day respiration to elevated atmospheric CO2. Plant Physiol. 2023, 191, 2204–2217. [Google Scholar] [CrossRef]
  39. Tjoelker, M.G.; Oleksyn, J.; Lorenc-Plucinska, G.; Reich, P.B. Acclimation of respiratory temperature responses in northern and southern populations of Pinus banksiana. New Phytol. 2009, 181, 218–229. [Google Scholar] [CrossRef] [PubMed]
  40. Li, X.; Zhang, G.Q.; Sun, B.; Zhang, S.; Zhang, Y.Q.; Liao, Y.W.K.; Zhou, Y.H.; Xia, X.J.; Shi, K.; Yu, J.Q. Stimulated Leaf Dark Respiration in Tomato in an Elevated Carbon Dioxide Atmosphere. Sci. Rep. 2013, 3, 3433. [Google Scholar] [CrossRef]
  41. Zhang, G.; Li, Y.; Li, F.; Xu, Z.; Sun, Y. Effects of root age on biomass and leaf nutrition in tetraploid Robinia pseudoacacia. J. Beijing For. Univ. 2009, 31, 37–41. [Google Scholar]
  42. Tan, X.; Peng, Z.; Jia, Z.; Ma, L. Influence of air temperatures on photosynthetic light-response curves of Robinia pseudoacacia L. J. Beijing For. Univ. 2010, 32, 64–68. [Google Scholar]
  43. Meng, F.J.; Pang, H.Y.; Huang, F.L.; Liu, L.; Wang, Y.J. Tetraploid black locust Robinia pseudoacacia L. Increased salt tolerance by activation of the antioxidant system. Biotechnol. Biotechnol. Equip. 2012, 26, 3351–3358. [Google Scholar] [CrossRef]
  44. Zhang, Y.; Luo, X.; Shen, Y. Dynamic changes of anti-oxidation system in new cultvars of Robinia pseudoacacia under gradual drought stress of soil. J. Zhejiang For. Coll. 2005, 22, 166–169. [Google Scholar]
  45. Liu, J.; Zhang, Z.; Li, Y.; Han, J.; Si, H.; Mi, Y.; Wang, S.; Wei, X.; Yang, H.; Sun, Y.; et al. Effects of the vegetative propagation method on juvenility in Robinia pseudoacacia L. For. Res. 2022, 2, 17. [Google Scholar] [CrossRef]
  46. Xu, F.; Jiang, M.; Meng, F. Short-term effect of elevated CO2 concentration (0.5%) on mitochondria in diploid and tetraploid black locust Robinia pseudoacacia L. Ecol. Evol. 2017, 7, 4651–4660. [Google Scholar] [CrossRef] [PubMed]
  47. AbdElgawad, H.; Zinta, G.; Hornbacher, J.; Papenbrock, J.; Markakis, M.N.; Asard, H.; Beemster, G.T.S. Elevated CO2 mitigates the impact of drought stress by upregulating glucosinolate metabolism in Arabidopsis thaliana. Plant Cell Environ. 2023, 46, 812–830. [Google Scholar] [CrossRef] [PubMed]
  48. Maurer, R.; Tapia, M.E.; Shor, A.C. Exogenous Root Uptake of Glycine Betaine Mitigates Improved Tolerance to Salinity Stress in Avicenna germinans under Ambient and Elevated CO2 Conditions. FASEB J. 2020, 34, 1–1. [Google Scholar] [CrossRef]
  49. Cohen, I.; Lichston, J.E.; de Macêdo, C.E.C.; Rachmilevitch, S. Leaf coordination between petiole vascular development and water demand in response to elevated CO2 in tomato plants. Plant Direct 2022, 6, e371. [Google Scholar] [CrossRef] [PubMed]
  50. Souza, J.P.; Melo, N.M.J.; Pereira, E.G.; Halfeld, A.D.; Gomes, I.N.; Prado, C. Responses of woody Cerrado species to rising atmospheric CO2 concentration and water stress: Gains and losses. Funct. Plant Biol. 2016, 43, 1183–1193. [Google Scholar] [CrossRef]
  51. Silvola, J.; Ahlholm, U. Photosynthesis in willows (Salix × dasyclados) grown at different CO2 concentrations and fertilization levels. Oecologia 1992, 91, 208–213. [Google Scholar] [CrossRef]
  52. Ellsworth, D.S.; Reich, P.B.; Naumburg, E.S.; Koch, G.W.; Kubiske, M.E.; Smith, S.D. Photosynthesis, carboxylation and leaf nitrogen responses of 16 species to elevated pCO2 across four free-air CO2 enrichment experiments in forest, grassland and desert. Glob. Chang. Biol. 2004, 10, 2121–2138. [Google Scholar] [CrossRef]
  53. Nowak, R.S.; Ellsworth, D.S.; Smith, S.D. Functional responses of plants to elevated atmospheric CO2—Do photosynthetic and productivity data from FACE experiments support early predictions? New Phytol. 2004, 162, 253–280. [Google Scholar] [CrossRef]
  54. Li, C.T.; Trigani, K.; Zuñiga, C.; Eng, R.; Chen, E.; Zengler, K.; Betenbaugh, M.J. Examining the impact of carbon dioxide levels and modulation of resulting hydrogen peroxide in Chlorella vulgaris. Algal Res.-Biomass Biofuels Bioprod. 2021, 60, 102492. [Google Scholar] [CrossRef]
  55. Ezraty, B.; Chabalier, M.; Ducret, A.; Maisonneuve, E.; Dukan, S. CO2 exacerbates oxygen toxicity. Embo Rep. 2011, 12, 321–326. [Google Scholar] [CrossRef] [PubMed]
  56. Gollan, P.J.; Tikkanen, M.; Aro, E.M. Photosynthetic light reactions: Integral to chloroplast retrograde signalling. Curr. Opin. Plant Biol. 2015, 27, 180–191. [Google Scholar] [CrossRef]
  57. Foyer, C.H. Reactive oxygen species, oxidative signaling and the regulation of photosynthesis. Environ. Exp. Bot. 2018, 154, 134–142. [Google Scholar] [CrossRef]
  58. Dinakar, C.; Vishwakarma, A.; Raghayendra, A.S.; Padmasree, K. Alternative Oxidase Pathway Optimizes Photosynthesis During Osmotic and Temperature Stress by Regulating Cellular ROS, Malate Valve and Antioxidative Systems. Front. Plant Sci. 2016, 7, 171347. [Google Scholar] [CrossRef]
  59. Mallick, N.; Mohn, F.H. Reactive oxygen species: Response of algal cells. J. Plant Physiol. 2000, 157, 183–193. [Google Scholar] [CrossRef]
  60. del Río, L.A.; Sandalio, L.M.; Corpas, F.J.; Palma, J.M.; Barroso, J.B. Reactive oxygen species and reactive nitrogen species in peroxisomes.: Production, scavenging, and role in cell signaling. Plant Physiol. 2006, 141, 330–335. [Google Scholar] [CrossRef]
  61. Asada, K. Production and scavenging of reactive oxygen species in chloroplasts and their functions. Plant Physiol. 2006, 141, 391–396. [Google Scholar] [CrossRef]
  62. Zhu, L.; Wu, J.; Li, M.; Fang, H.; Zhang, J.; Chen, Y.; Chen, J.; Cheng, T. Genome-wide discovery of CBL genes in Nitraria tangutorum Bobr. and functional analysis of NtCBL1-1 under drought and salt stress. For. Res. 2023, 3, 28. [Google Scholar] [CrossRef]
  63. Chen, Y.; Li, M.; Gong, L.; Song, Y. Effects of Exogenous Signal Substances on POD, CAT Activity and ROS Content in Cistanche deserticola Y. C. Ma Seeds During Germination and Haustorium Formation. Chin. Agric. Sci. Bull. 2014, 30, 128–132. [Google Scholar]
  64. Li, Y.; Zhou, Q.; Li, F.; Liu, X.; Luo, Y. Effects of tetrabromobisphenol A as an emerging pollutant on wheat (Triticum aestivum) at biochemical levels. Chemosphere 2008, 74, 119–124. [Google Scholar] [CrossRef] [PubMed]
  65. Ge, H.; Zhang, F. Effects of Tetrabromobisphenol a Stress on Growth and Physiological Characteristics of Soybean Seedling. Bull. Environ. Contam. Toxicol. 2017, 98, 141–146. [Google Scholar] [CrossRef] [PubMed]
  66. Zhang, Y.; Chen, Y.; Guo, Y.; Ma, Y.; Yang, M.; Fu, R.; Sun, Y. Elevated CO2 delayed yellowing by maintaining chlorophyll biosynthesis and inhibiting chlorophyll degradation and carotenoid accumulation of postharvest broccoli. Postharvest Biol. Technol. 2022, 194, 112089. [Google Scholar] [CrossRef]
  67. Steer, B.T.; Walker, D.A. Inhibition of Chlorophyll Synthesis by High Concentrations of Carbon Dioxide. Plant Physiol. 1965, 40, 577–581. [Google Scholar] [CrossRef] [PubMed]
  68. Li, Y.L.; Jin, Z.X.; Li, J.M. Interactive effects of copper stress and arbuscular mycorrhizal fungi on photosynthetic characteristics and chlorophyll fluorescence parameters of elsholtzia splendens. Pak. J. Bot. 2017, 49, 1531–1540. [Google Scholar]
  69. Laisk, A.; Oja, V.; Rasulov, B.; Eichelmann, H.; Sumberg, A. Quantum Yields and Rate Constants of Photochemical and Nonphotochemical Excitation Quenching (Experiment and Model). Plant Physiol. 1997, 115, 803–815. [Google Scholar] [CrossRef] [PubMed]
  70. Dahal, K.; Vanlerberghe, G.C. Growth at Elevated CO2 Requires Acclimation of the Respiratory Chain to Support Photosynthesis. Plant Physiol. 2018, 178, 82–100. [Google Scholar] [CrossRef]
  71. Gandin, A.; Duffes, C.; Day, D.A.; Cousins, A.B. The Absence of Alternative Oxidase AOX1A Results in Altered Response of Photosynthetic Carbon Assimilation to Increasing CO2 in Arabidopsis thaliana. Plant Cell Physiol. 2012, 53, 1627–1637. [Google Scholar] [CrossRef]
  72. Hedrich, R.; Neimanis, S.; Savchenko, G.; Felle, H.H.; Kaiser, W.M.; Heber, U. Changes in apoplastic pH and membrane potential in leaves in relation to stomatal responses to CO2, malate, abscisic acid or interruption of water supply. Planta 2001, 213, 594–601. [Google Scholar] [CrossRef] [PubMed]
  73. Engineer, C.B.; Hashimoto-Sugimoto, M.; Negi, J.; Israelsson-Nordström, M.; Azoulay-Shemer, T.; Rappel, W.J.; Iba, K.; Schroeder, J.I. CO2 Sensing and CO2 peculation of Stomatal Conductance: Advances and Open Questions. Trends Plant Sci. 2016, 21, 16–30. [Google Scholar] [CrossRef] [PubMed]
  74. Mao, Z.J.; Wang, Y.J.; Wang, X.W.; Voronin, P.Y. Effect of doubled CO2 on morphology: Inhibition of stomata development in growing birch Betula platyphylla Suk. leaves. Russ. J. Plant Physiol. 2005, 52, 171–175. [Google Scholar] [CrossRef]
  75. Ma, X.N.; Bai, L. Elevated CO2 and Reactive Oxygen Species in Stomatal Closure. Plants 2021, 10, 410. [Google Scholar] [CrossRef] [PubMed]
  76. Wang, Z.; Li, L.; Khan, D.; Chen, Y.; Pu, X.; Wang, X.; Guan, M.; Rengel, Z.; Chen, Q. Nitric oxide acts downstream of reactive oxygen species in phytomelatonin receptor 1 (PMTR1)-mediated stomatal closure in Arabidopsis. J. Plant Physiol. 2023, 282, 153917. [Google Scholar] [CrossRef] [PubMed]
  77. Correa Rojo, A.; Heylen, D.; Aerts, J.; Thas, O.; Hooyberghs, J.; Ertaylan, G.; Valkenborg, D. Towards Building a Quantitative Proteomics Toolbox in Precision Medicine: A Mini-Review. Front. Physiol. 2021, 12, 723510. [Google Scholar] [CrossRef]
  78. Wang, L.; Wang, F.; Zhang, Y.; Ma, Y.; Guo, Y.; Zhang, X. Enhancing the ascorbate–glutathione cycle reduced fermentation by increasing NAD+ levels during broccoli head storage under controlled atmosphere. Postharvest Biol. Technol. 2020, 165, 111169. [Google Scholar] [CrossRef]
  79. Sicher, R.C.; Hatch, A.L.; Stumpf, D.K.; Jensen, R.G. Ribulose 1,5-Bisphosphate and Activation of the Carboxylase in the Chloroplast. Plant Physiol. 1981, 68, 252–255. [Google Scholar] [CrossRef]
  80. Fish, L.E.; Bogorad, L. Identification and analysis of the maize P700 chlorophyll a apoproteins PSI-A1 and PSI-A2 by high pressure liquid chromatography analysis and partial sequence determination. J. Biol. Chem. 1986, 261, 8134–8139. [Google Scholar] [CrossRef]
  81. Fernandez-Velasco, J.G.; Jamshidi, A.; Gong, X.S.; Zhou, J.; Ueng, R.Y. Photosynthetic electron transfer through the cytochrome b6f complex can bypass cytochrome f. J. Biol. Chem. 2001, 276, 30598–30607. [Google Scholar] [CrossRef]
  82. Ustynyuk, L.Y.; Tikhonov, A.N. Plastoquinol Oxidation: Rate-Limiting Stage in the Electron Transport Chain of Chloroplasts. Biochemistry 2022, 87, 1084–1097. [Google Scholar] [CrossRef] [PubMed]
  83. de Weerd, F.L.; van Stokkum, I.H.M.; van Amerongen, H.; Dekker, J.P.; van Grondelle, R. Pathways for energy transfer in the core light-harvesting complexes CP43 and CP47 of photosystem II. Biophys. J. 2002, 82, 1586–1597. [Google Scholar] [CrossRef]
  84. Bricker, T.M.; Frankel, L.K. The structure and function of CP47 and CP43 in Photosystem II. Photosynth. Res. 2002, 72, 131–146. [Google Scholar] [CrossRef] [PubMed]
  85. Karapetyan, N.V. Organization and role of the long-wave chlorophylls in the photosystem I of the Cyanobacterium spirulina. Membr. Cell Biol. 1998, 12, 571–584. [Google Scholar] [PubMed]
  86. Ishikita, H.; Loll, B.; Biesiadka, J.; Saenger, W.; Knapp, E.W. Redox potentials of chlorophylls in the photosystem II reaction center. Biochemistry 2005, 44, 4118–4124. [Google Scholar] [CrossRef]
  87. Xu, Y.H.; Liu, R.; Yan, L.; Liu, Z.Q.; Jiang, S.C.; Shen, Y.Y.; Wang, X.F.; Zhang, D.P. Light-harvesting chlorophyll-binding proteins are required for stomatal response to abscisic acid in Arabidopsis. J. Exp. Bot. 2012, 63, 1095–1106. [Google Scholar] [CrossRef] [PubMed]
  88. Brettel, K.; Leibl, W. Electron transfer in photosystem I. Biochim. Biophys. Acta 2001, 1507, 100–114. [Google Scholar] [CrossRef] [PubMed]
  89. Sunde, R.A.; Hoekstra, W.G. Structure, Synthesis and Function of Glutathione Peroxidase. Nutr. Rev. 1980, 38, 265–273. [Google Scholar] [CrossRef]
  90. Babich, H.; Liebling, E.J.; Burger, R.F.; Zuckerbraun, H.L.; Schuck, A.G. Choice of DMEM, formulated with or without pyruvate, plays an important role in assessing the in vitro cytotoxicity of oxidants and prooxidant nutraceuticals. Vitr. Cell. Dev. Biol. -Anim. 2009, 45, 226–233. [Google Scholar] [CrossRef]
  91. Zhang, D.; Li, S.; Wang, G.; Guo, D.; Xing, K.; Zhang, S. Biochemical responses of the copepod Centropages tenuiremis to CO2-driven acidified seawater. Water Sci. Technol. 2012, 65, 30–37. [Google Scholar] [CrossRef]
  92. Liu, Q.; Wang, S.W.; Du, Y.L.; Yin, K.D. Improved drought tolerance in soybean by protein elicitor AMEP412 induced ROS accumulation and scavenging. Biotechnol. Biotechnol. Equip. 2022, 36, 401–412. [Google Scholar] [CrossRef]
  93. Calatayud, A.; San Bautista, A.; Pascual, B.; Maroto, J.V.; Lopez-Galarza, S. Use of chlorophyll fluorescence imaging as diagnostic technique to predict compatibility in melon graft. Sci. Hortic. 2013, 149, 13–18. [Google Scholar] [CrossRef]
  94. Xu, L.; Hu, Y.; Jin, G.; Lei, P.; Sang, L.; Luo, Q.; Liu, Z.; Guan, F.; Meng, F.; Zhao, X. Physiological and Proteomic Responses to Drought in Leaves of Amygdalus mira (Koehne) Yü et Lu. Front. Plant Sci. 2021, 12, 620499. [Google Scholar] [CrossRef]
  95. Perez-Riverol, Y.; Bai, J.W.; Bandla, C.; García-Seisdedos, D.; Hewapathirana, S.; Kamatchinathan, S.; Kundu, D.J.; Prakash, A.; Frericks-Zipper, A.; Eisenacher, M.; et al. The PRIDE database resources in 2022: A hub for mass spectrometry-based proteomics evidences. Nucleic Acids Res. 2022, 50, D543–D552. [Google Scholar] [CrossRef]
Figure 1. Physiological response of the tetraploid R. pseudoacacia to high CO2 concentrations. (a) Phenotypes of R. pseudoacacia under CO2 treatment. Plants were exposed to CO2 stress for 9 days and then recovered to air CO2 levels for 3 days. CK-H2O: control; CK-PA: pyruvic acid-treated; CK-SHAM: salicylhydroxamic acid-treated; T-H2O: CO2-treated; T-PA: co-treated with CO2 and PA; and T-SHAM: co-treated with CO2 and SHAM. Bar: 20 cm. Effect of high CO2 concentrations on relative electrical conductivity (b), relative water content (c), net photosynthetic rate (Pn) (d), stomatal conductance (Gs) (e), intercellular CO2 concentration (Ci) (f), and transpiration rate (Tr) (g). Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Figure 1. Physiological response of the tetraploid R. pseudoacacia to high CO2 concentrations. (a) Phenotypes of R. pseudoacacia under CO2 treatment. Plants were exposed to CO2 stress for 9 days and then recovered to air CO2 levels for 3 days. CK-H2O: control; CK-PA: pyruvic acid-treated; CK-SHAM: salicylhydroxamic acid-treated; T-H2O: CO2-treated; T-PA: co-treated with CO2 and PA; and T-SHAM: co-treated with CO2 and SHAM. Bar: 20 cm. Effect of high CO2 concentrations on relative electrical conductivity (b), relative water content (c), net photosynthetic rate (Pn) (d), stomatal conductance (Gs) (e), intercellular CO2 concentration (Ci) (f), and transpiration rate (Tr) (g). Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Ijms 25 05262 g001
Figure 2. Effect of high CO2 on NBT staining of the fresh leaves (a,b), O2·− level (c), DAB staining of the fresh leaves (d,e), H2O2 content (f), SOD (g), CAT (h), GR (i), and APX (j). Bar: 1 cm. Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Figure 2. Effect of high CO2 on NBT staining of the fresh leaves (a,b), O2·− level (c), DAB staining of the fresh leaves (d,e), H2O2 content (f), SOD (g), CAT (h), GR (i), and APX (j). Bar: 1 cm. Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Ijms 25 05262 g002
Figure 3. Effect of high CO2 concentration on chlorophyll content and chlorophyll fluorescence parameters in R. pseudoacacia. Chl a content (a), Chl b content (b), Chl a/b (c), minimal fluorescence (Fo) (d), maximal fluorescence (Fm) (e), primary light energy conversion efficiency of PSII (Fv/Fm) (f), nonphotochemical quenching (NPQ) (g), photochemical quenching (qP) (h), and relative electron transport rate (ETR) (i). Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Figure 3. Effect of high CO2 concentration on chlorophyll content and chlorophyll fluorescence parameters in R. pseudoacacia. Chl a content (a), Chl b content (b), Chl a/b (c), minimal fluorescence (Fo) (d), maximal fluorescence (Fm) (e), primary light energy conversion efficiency of PSII (Fv/Fm) (f), nonphotochemical quenching (NPQ) (g), photochemical quenching (qP) (h), and relative electron transport rate (ETR) (i). Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Ijms 25 05262 g003
Figure 4. Effect of high CO2 concentration on respiratory parameters and mitochondrial electron transport chain complex enzyme activities in R. pseudoacacia. Cytochrome pathway capacity (Vcyt) (a), alternative pathway capacity (Valt) (b), total respiration (Vt) (c), AOX enzyme activity (d), complex I enzyme activity (e), complex II enzyme activity (f), complex III enzyme activity (g), and complex IV enzyme activity (h). Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Figure 4. Effect of high CO2 concentration on respiratory parameters and mitochondrial electron transport chain complex enzyme activities in R. pseudoacacia. Cytochrome pathway capacity (Vcyt) (a), alternative pathway capacity (Valt) (b), total respiration (Vt) (c), AOX enzyme activity (d), complex I enzyme activity (e), complex II enzyme activity (f), complex III enzyme activity (g), and complex IV enzyme activity (h). Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Ijms 25 05262 g004
Figure 5. Effects of high CO2 concentration on stomatal morphology (a), stomatal length (b), stomatal width (c), and stomatal aperture of plant (d). Bar: 50 μm. Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Figure 5. Effects of high CO2 concentration on stomatal morphology (a), stomatal length (b), stomatal width (c), and stomatal aperture of plant (d). Bar: 50 μm. Six biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Ijms 25 05262 g005
Figure 6. 2′,7′-Dichlorodihydrofluorescein diacetate (H2DCF-DA) (a,c) and 5,6-Diaminofluorescein diacetate (DAF-2DA) (b,d) staining of guard cells under high CO2 conditions. Ten biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Figure 6. 2′,7′-Dichlorodihydrofluorescein diacetate (H2DCF-DA) (a,c) and 5,6-Diaminofluorescein diacetate (DAF-2DA) (b,d) staining of guard cells under high CO2 conditions. Ten biological replicates were analyzed, and the error bars represent the SE. Asterisks indicate significant differences as determined by Dunnett’s test. (* p ≤ 0.05; ** p ≤ 0.01; *** p ≤ 0.001).
Ijms 25 05262 g006
Figure 7. Volcano maps of differentially expressed proteins (DEPs) in R. pseudoacacia under CO2 treatment. The horizontal coordinate is log2 (fold change), and the vertical coordinate is the negative logarithm of the p-value of the t-test significance test –log10 (padj). (a) T-H2O vs. CK-H2O; (b) CK-PA vs. CK-H2O; (c) CK-SHAM vs. CK-H2O; (d) T-PA vs. CK-H2O; (e) T-SHAM vs. CK-H2O; (f) T-PA vs. T-H2O; and (g) T-SHAM vs. T-H2O. Red represents upregulated proteins, blue represents downregulated proteins, grey represents proteins with no differential change, and the dotted grey line in the middle represents the threshold line for the DAP screening criteria.
Figure 7. Volcano maps of differentially expressed proteins (DEPs) in R. pseudoacacia under CO2 treatment. The horizontal coordinate is log2 (fold change), and the vertical coordinate is the negative logarithm of the p-value of the t-test significance test –log10 (padj). (a) T-H2O vs. CK-H2O; (b) CK-PA vs. CK-H2O; (c) CK-SHAM vs. CK-H2O; (d) T-PA vs. CK-H2O; (e) T-SHAM vs. CK-H2O; (f) T-PA vs. T-H2O; and (g) T-SHAM vs. T-H2O. Red represents upregulated proteins, blue represents downregulated proteins, grey represents proteins with no differential change, and the dotted grey line in the middle represents the threshold line for the DAP screening criteria.
Ijms 25 05262 g007
Figure 8. Cluster analysis of orthologous groups (KOG/COG)of DEPs in R. pseudoacacia under CO2 treatment. (a) T-H2O vs. CK-H2O; (b) CK-PA vs. CK-H2O; (c) CK-SHAM vs. CK-H2O; (d) T-PA vs. CK-H2O; (e) T-SHAM vs. CK-H2O; (f) T-PA vs. T-H2O; and (g) T-SHAM vs. T-H2O.
Figure 8. Cluster analysis of orthologous groups (KOG/COG)of DEPs in R. pseudoacacia under CO2 treatment. (a) T-H2O vs. CK-H2O; (b) CK-PA vs. CK-H2O; (c) CK-SHAM vs. CK-H2O; (d) T-PA vs. CK-H2O; (e) T-SHAM vs. CK-H2O; (f) T-PA vs. T-H2O; and (g) T-SHAM vs. T-H2O.
Ijms 25 05262 g008
Figure 9. KEGG enrichment analysis of DEPs identified in R. pseudoacacia under CO2 treatment. (a) T-H2O vs. CK-H2O; (b) CK-PA vs. CK-H2O; (c) CK-SHAM vs. CK-H2O; (d) T-PA vs. CK-H2O; (e) T-SHAM vs. CK-H2O; (f) T-PA vs. T-H2O; and (g) T-SHAM vs. T-H2O. The x-axis represents the enrichment factor, and the y-axis represents the pathway of enrichment. Larger orange points represent major pathway enrichment and higher pathway impact values, respectively.
Figure 9. KEGG enrichment analysis of DEPs identified in R. pseudoacacia under CO2 treatment. (a) T-H2O vs. CK-H2O; (b) CK-PA vs. CK-H2O; (c) CK-SHAM vs. CK-H2O; (d) T-PA vs. CK-H2O; (e) T-SHAM vs. CK-H2O; (f) T-PA vs. T-H2O; and (g) T-SHAM vs. T-H2O. The x-axis represents the enrichment factor, and the y-axis represents the pathway of enrichment. Larger orange points represent major pathway enrichment and higher pathway impact values, respectively.
Ijms 25 05262 g009
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

Li, J.; Zhang, S.; Lei, P.; Guo, L.; Zhao, X.; Meng, F. Physiological and Proteomic Responses of the Tetraploid Robinia pseudoacacia L. to High CO2 Levels. Int. J. Mol. Sci. 2024, 25, 5262. https://doi.org/10.3390/ijms25105262

AMA Style

Li J, Zhang S, Lei P, Guo L, Zhao X, Meng F. Physiological and Proteomic Responses of the Tetraploid Robinia pseudoacacia L. to High CO2 Levels. International Journal of Molecular Sciences. 2024; 25(10):5262. https://doi.org/10.3390/ijms25105262

Chicago/Turabian Style

Li, Jianxin, Subin Zhang, Pei Lei, Liyong Guo, Xiyang Zhao, and Fanjuan Meng. 2024. "Physiological and Proteomic Responses of the Tetraploid Robinia pseudoacacia L. to High CO2 Levels" International Journal of Molecular Sciences 25, no. 10: 5262. https://doi.org/10.3390/ijms25105262

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