Next Article in Journal
Age- and Brain Region-Specific Changes of Glucose Metabolic Disorder, Learning, and Memory Dysfunction in Early Alzheimer’s Disease Assessed in APP/PS1 Transgenic Mice Using 18F-FDG-PET
Next Article in Special Issue
Comparative Proteomic and Physiological Analysis Reveals the Variation Mechanisms of Leaf Coloration and Carbon Fixation in a Xantha Mutant of Ginkgo biloba L.
Previous Article in Journal
Dissecting Candida albicans Infection from the Perspective of C. albicans Virulence and Omics Approaches on Host–Pathogen Interaction: A Review
Previous Article in Special Issue
Proteomic Analysis of Hylocereus polyrhizus Reveals Metabolic Pathway Changes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Drought-Responsive Mechanisms in Plant Leaves Revealed by Proteomics

Development Centre of Plant Germplasm Resources, College of Life and Environmental Sciences, Shanghai Normal University, Shanghai 200234, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2016, 17(10), 1706; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17101706
Submission received: 30 June 2016 / Revised: 6 September 2016 / Accepted: 22 September 2016 / Published: 18 October 2016
(This article belongs to the Special Issue Plant Proteomic Research)

Abstract

:
Plant drought tolerance is a complex trait that requires a global view to understand its underlying mechanism. The proteomic aspects of plant drought response have been extensively investigated in model plants, crops and wood plants. In this review, we summarize recent proteomic studies on drought response in leaves to reveal the common and specialized drought-responsive mechanisms in different plants. Although drought-responsive proteins exhibit various patterns depending on plant species, genotypes and stress intensity, proteomic analyses show that dominant changes occurred in sensing and signal transduction, reactive oxygen species scavenging, osmotic regulation, gene expression, protein synthesis/turnover, cell structure modulation, as well as carbohydrate and energy metabolism. In combination with physiological and molecular results, proteomic studies in leaves have helped to discover some potential proteins and/or metabolic pathways for drought tolerance. These findings provide new clues for understanding the molecular basis of plant drought tolerance.

1. Introduction

Drought is an inevitable and recurring feature of world climate. Despite our efforts to forecast its onset and modify its impact, drought remains the single most important factor affecting worldwide crop growth and productivity [1]. More importantly, although plant drought response has been extensively studied [2,3,4], there are still no economically practical technological means to facilitate crop production under drought [5]. Therefore, there is an urgent need to further understand and enhance crop tolerance to drought stress.
Drought stress induces a number of changes at the morphological, physiological and biochemical level in all plant organs [6]. Plants have evolved several strategies to cope with drought stress, including drought escape via a short life cycle or developmental plasticity, drought avoidance via enhanced water uptake and reduced water loss, as well as drought tolerance via osmotic adjustment, antioxidant capacity, and desiccation tolerance [7]. With the development of high-throughput sequencing and various omic technologies, large amounts of sequence data sets and global changes in gene expressions have been reported. For example, using genomics and quantitative trait loci (QTL) mapping approaches, a number of major genes or QTL for drought stress resistance were identified in Arabidopsis [8], pearl millet (Pennisetum glaucum) [9], Festuca pratensis (Huds.) [10], wheat (Triticum aestivum) [11], maize (Zea mays) [12], and common bean (Phaseolus vulgaris) [13]. Besides, microarray analysis of drought response in shoots of two T. aestivum cultivars (i.e., TAM 111 and TAM 112) have revealed that 1657 transcripts were commonly altered in both cultivars, but 474 and 1540 transcripts were unique to each cultivar, respectively [14]. In addition, genome-wide transcriptional analysis of Populus euphratica under four different drought treatments (watering with 70%, 50%, 20%, and 5% of daily evaporation for seven weeks) have identified 952, 1354, 2138 and 2360 transcripts, respectively. Among them, some candidate drought-responsive genes have a high potential to be used for crop breeding with drought tolerance. Those genes are involved in the biosynthesis of plant hormones, various signaling pathways (e.g., ZEP, PYL, PP2C, SnRK2, ACO, ACS, ETR1, and ETO1), osmoprotective pathways (e.g., BADH, P5CS, PDH1, TPS1, and LEA), as well as the detoxification of reactive oxygen species (e.g., APX, SOD, GR, and ALDH7) [15].
However, the mRNA levels usually do not correlate well with the protein abundances and functions, due to various post-translational modifications. High-throughput proteomics has proved to be a powerful tool for the comprehensive identification of drought-responsive proteins in plants [16,17]. In previous investigations, more than 2200 drought-responsive protein species have been identified in leaves from 25 plant species, including 18 herbs, 3 shrubs and 4 trees (Table 1, Supplementary Table S1). In the present review, these drought-responsive protein species are defined as 440 unique proteins on the basis of their protein sequence homology and functional domain similarity. These proteins are mainly involved in signaling, transcription, stress and defense, protein synthesis, folding and degradation, photosynthesis and photorespiration, carbohydrate and energy metabolism, membrane and transport, cell structure and cell cycle, nitrogen assimilation and amino acid metabolism, as well as fatty acid metabolism (Figure 1). In addition, we also found some drought-responsive phosphoproteins in four phosphoproteomics studies (Table 1, Supplementary Table S2). Phosphorylation is one of the most important post-translational modifications (PTMs) that modulates protein activity, protein–protein interaction and cellular localization. These drought-responsive phosphoproteins were mainly involved in signaling, transcription, photosynthesis and carbon metabolism, as well as in protein synthesis and turnover. The integrative analysis of physiological, molecular, and proteomic characteristics provides new clues for further understanding plant drought tolerance.

2. Drought Sensing and Signaling

2.1. Possible Drought Receptor

Although no specific receptors for drought sensing have been found in plants, a drought-responsive photoreceptor, phytochrome C1, was identified in Z. mays [41] (Figure 2A). Phytochrome is widely believed to regulate the transcription of light-responsive genes by modulating the activity of several transcription factors under both biotic and abiotic stresses [65]. In Arabidopsis, three phytochrome genes (i.e., PHYA, PHYB, and PHYE) are involved in suppressing drought tolerance [66]. These results imply a possible function of phytochrome C in mediating osmotic stress.

2.2. G Proteins

G proteins constitute one of the most important cell signaling cascades. Proteomic studies revealed that two G protein subunits (alpha subunit and beta subunit), several small G proteins (e.g., Ras-related protein Rab7 and Ras-related nuclear protein Ran), and a Ran-binding protein 1 were increased in drought-treated leaves (Supplementary Table S1, Figure 2A). Their functions in drought tolerance have been previously reported. The G protein alpha subunit may play a positive role in regulation of drought stress [67,68], whereas the G protein beta subunit in Arabidopsis may negatively regulate drought tolerance [69,70]. Besides, Rab proteins are the largest branch of the small G protein superfamily, which are involved in vesicle trafficking, intracellular signaling events and many important physiological processes, such as polar growth, plant hormone signal cross-talk, and stress response [71]. One of the Rab proteins, Rab 7, has been suggested to be involved in drought stress tolerance. The PgRab7 gene was upregulated by dehydration in P. glaucum [72], while overexpression of the peanut AhRabG3f exhibited an enhanced tolerance to drought stress in transgenic peanut (Arachis hypogaea L.) [73]. Additionally, two isoforms of GDP dissociation inhibitor (GDI), involved in the regulation of Rab family activity, were increased in drought-stressed leaves of Oryza sativa [29]. The expression of the MiRab-GDI gene was induced in Mangifera indica under drought stress conditions [74], but the exact function of GDI in response to drought stress remains to be elucidated. In addition, Ran is another member of the small G protein group, which is involved in nucleo-cytoplasmic transportation of proteins and RNA, the formation of spindle asters, and the reassembly of the nuclear envelope in mitotic cells [71]. Ran-binding protein 1 is the major effector of Ran. The significant increase of Ran in Poa pratensis [23] and Ran-binding protein 1 in Z. mays under drought stress conditions [41] indicated the important role of cell cycle and DNA synthesis in drought tolerance. However, the drought-decreased Ran/TC4 in Cistus albidus under no irrigation for 107 days [53] implies their function is dependent on the drought intensity.

2.3. Ca2+ Signaling and Protein Kinase

Under drought stress, calcium acts as a second messenger, which is employed to regulate specific protein kinase activity and downstream gene expression [75]. In proteomic studies, the abundances of several calcium binding proteins (CaBs), such as calmodulin (CaM), calcium sensing receptor (CaSR), calreticulin (CRT), and calcium-dependent protein kinase (CDPK), were changed in response to drought (Figure 2A). Among them, CRT is a unique endoplasmic reticulum (ER) luminal Ca2+ binding chaperone, which plays a role in many cellular processes [76]. The increase of CRT abundance enhanced the survival of T. aestivum plants under drought condition [77], and the TaCRT-overexpressing tobacco (Nicotiana benthamiana) plants exhibited enhanced drought resistance [78]. Proteomic studies revealed a drought-induced CRT1 in Glycine max [17], but a drought-reduced CRT1 in Quercus robur under prolonged (65 days) drought stress [61]. These results indicate that CRT functions in drought tolerance in a stress intensity-dependent manner. Besides, CDPKs regulate the downstream components in calcium-mediated signal transduction. In proteomic studies, the protein abundance and phosphorylation level of CDPKs were generally decreased in plants under drought. However, the abundance of CDPKs increased in the drought-tolerant genotype of Z. mays [41]. It has been reported that overexpression of the CDPK gene can enhance drought tolerance in transgenic Arabidopsis [79,80] and O. sativa [81]. These reports imply the important role of CDPKs in plant stress tolerance. The phosphorylation states of several protein kinases (e.g., serine/threonine-protein kinase, germinal center kinase (GCK)-like kinase MIK, receptor-like protein kinase HERK 1-like, phototropin family protein kinase, and salt-inducible protein kinase) were also changed in response to drought stress, implying their regulation in the drought response signaling pathway. In addition, the dephosphorylation mediated by protein phosphatases is an important event in the signal transduction process that regulates various cellular activities [82]. In proteomic studies, the phosphorylation level of protein phosphatase 2C (PP2C) and phospholipase D (PLD) in Z. mays [41] were drought-increased (Figure 2A). PP2C is known to be a negative regulator for plant drought tolerance in the abscisic acid (ABA) signaling pathway, which can inhibit the activity of SnRK, leading to a decrease of the phosphorylation of its substrates in the signaling cascade [83,84,85]. Both phosphorylation level [40,44] and abundance of PP2C [41,86] were obviously affected by drought stress. These data suggest that the PP2C-involved ABA signaling pathway is crucial for drought response. Additionally, a dual function of PLD in plant drought response was reported, for example, in drought-sensitive cultivars, the activity and expression of PLD increased more obviously than in drought-tolerant ones [87,88], whereas overexpression of the PLDα gene can significantly enhance drought tolerance in transgenic Arabidopsis [89] and Populus tomentosa [90]. In addition, a high level of PLD promotes stomatal closure at earlier stages, but disrupts membranes in prolonged drought stress [91]. These data imply that, in drought, PLD functions in a condition-dependent manner.

2.4. 14-3-3 Proteins

In proteomic studies, some members of the 14-3-3 protein family and 14-3-3-like proteins were increased in drought-treated leaves from many plant species (Figure 2A, Supplementary Table S1). The 14-3-3 proteins are intracellular dimeric phosphoserine/threonine binding molecules that participate in a wide range of vital regulatory processes, including signaling, transmembrane receptors, and transcription activation [92]. 14-3-3 proteins are also known as positive regulators of H+-ATPase activity to control the electrochemical gradient across the plasma membrane, which contributes to the initiation of stress responses and other signal transduction pathways [93]. It has been reported that drought stress can directly alter the abundance of 14-3-3 proteins [94,95,96]. In addition, overexpression or silencing of the 14-3-3 protein genes can modulate drought tolerance of transgenic plants (e.g., Gossypium hirsutum and Arabidopsis) [97,98]. These results imply that 14-3-3 proteins may have diverse regulatory roles in leaves under drought stress, and the involvement of each 14-3-3 protein in drought-response needs to be further elucidated.

2.5. Ethylene and Auxin Signaling Pathways

It is well known that ABA is a key phytohormone that mediates the adaptive response to drought stress. Besides, other hormonal signals, such as ethylene, gibberellic acid, and jasmonic acid are also important cellular regulators in signal transduction pathway under drought conditions [99,100,101]. Proteomic studies revealed a drought-increased ethylene-responsive transcription factor (ERF) in Gossypium herbaceum [51] and some members of drought-responsive auxin-binding protein (ABP) family in Q. robur [61], Z. mays [41], and polar clones [59]. The ERF gene was found to be induced in G. herbaceum under drought stress [102,103] (Figure 2A). Overexpression of ERFs in various plants, such as sugarcane SodERF3 overexpression in tobacco, tomato TERF1 in rice, and Brassica rapa BrERF4 in Arabidopsis, can improve plant drought tolerance. However, careful attention should be taken in the definition of ERF roles in plant drought tolerance. Different ERF members may play different roles in plants in response to drought stress. For example, BpERF11 was found to negatively regulate osmotic tolerance in Betula platyphylla [104]. In addition, ABP functions as an auxin receptor, being involved in many development processes and drought response. However, as far as we know, little information is available for ABP members (i.e., ABP2, ABP20, and ABP19a) in response to drought stress. Additionally, a drought-increased TGF-β-receptor interacting protein 1 (TRIP1) was found in Sporobolus stapfianus [24]. TRIP-1 was phosphorylated by the brassinosteroid (BR)-insensitive I (BRI-1) protein, which is a serine/threonine kinase receptor essential for BR perception and signal transduction. The increase of TRIP-1 in S. stapfianus suggests that the BR signaling pathway would be triggered in response to water deficit.

3. Drought-Responsive Gene Expression Regulation

Drought-responsive gene expression is crucial in the transcriptional regulatory network [105]. Proteomic studies have found some gene expression-related proteins in the complex regulatory networks in leaves (Figure 2B). Chromatin structure modification is a prerequisite to sustain the transcriptional regulation which is involved in cell cycle progression. Significant abundance changes were observed in two chromatin structure modification-related proteins, which were histones and high mobility group proteins (HMG). Histones are the major proteins of chromatin, and the dynamic association of histones and their variants can regulate gene expression [106]. In proteomic studies, several histones (e.g., H1 and H2B) appeared to cause diverse abundance changes in different plant species in response to drought stress. For example, H2Bs were decreased in C. albidus [53] and Brassica napus [47], while histone H1 was decreased in a drought-sensitive Z. mays cultivar, but increased in a drought-tolerant one [41]. Similarly, the transcript and protein of histone H1 variant were all induced specifically in the tolerant genotype of G. herbaceum [107]. Besides, HMG is a highly conserved nuclear DNA-binding protein, which functions in DNA repair and chromatin modification after DNA damage [108]. A proteomic study revealed that two isoforms of HMG were drought-increased in drought-tolerant maize cultivar, but decreased in a sensitive cultivar [41]. Interestingly, the phosphorylation level of HMG was significantly decreased in a drought-tolerant wheat cultivar, but increased in a drought-sensitive one [40]. Previous studies have reported that phosphorylation of HMG reduced its binding to DNA, inhibiting replication and transcription [109]. These data imply that PTM (e.g., phosphorylation) of HMG is crucial for its function in plant drought tolerance.
In addition, RNA processing was also critical for plants to cope with drought stress. The abundances of several RNA processing-related proteins were changed in plants under drought stress (Figure 2C). Among them, five glycine-rich RNA binding proteins (GR-RBPs) were drought-increased, while three GR-RBPs were drought-decreased. RNA binding proteins (RBPs) can bind to RNA molecules, which are involved in almost all aspects of post-transcriptional gene regulation. The GR-RBP genes have been noted to be upregulated in response to water stress [110], which was suspected to function in the regulation of specific gene expression in response to stress. For example, the expression of GR-RBP gene in transgenic rice has been shown to have much higher recovery rates and grain yields when compared with that in wild-type plants under drought conditions [111]. However, the transgenic Arabidopsis and camellia plants that had an overexpressed GR-RBPa gene from Camelina sativa appeared to have a reduced drought tolerance [112]. Besides, proteomic studies revealed that the abundance and phosphorylation level of S-like ribonucleases (RNases) were significantly increased in rice under drought stress [30,34]. S-like RNases have acquired specialized functions such as stress regulation, defense against microorganisms, phosphate scavenging, and even nitrogen storage [113]. It has been reported that S-like RNases participated in the responses to salt, polyethylene glycol (PEG), and ABA [113]. All these have attracted great attention as researchers have attempted to further confirm their post-transcriptional regulation in response to drought stress. Additionally, an intron splicing related protein, and maturase K (MatK) appeared to show a drought decrease in poplar [60], and multiple organelle RNA editing factor 9, involved in RNA editing in mitochondria and plastids, was increased firstly, and then decreased in B. napus with the extension of drought stress [47]. Both of them are RNA processing-related proteins. Their changes indicated that the transcription regulation is diverse and complicates the study of plants’ ability to cope with drought stress.

4. Drought-Responsive Protein Synthesis and Turnover

Protein synthesis and turnover is one of the fundamental metabolic processes for plants to cope with drought stress. Proteomic investigations revealed that 16% of the drought-responsive proteins in leaves are attributed to protein synthesis and turnover functions (Figure 1). Several proteins are involved in protein biosynthesis, such as ribosomal protein (RP), elongation factor (EF), translation initiation factor (TIF), tRNA synthase (TRS), and ribosome recycling factor (RRF) (Figure 2D). Most of them exhibited an increase under drought stress, which would be beneficial for protein synthesis in response to specific drought conditions. Besides, the proteins functioning in protein folding and processing showed diverse changes among different plant species and cultivars (Figure 2E). For example, peptidyl-prolyl cis-trans isomerases (PPIases) were significantly increased in G. max [17], T. aestivum [35], O. sativa [29], and Q. robur [61], but decreased in a drought-sensitive cultivar of Phaseolus vulgaris [49]. Protein disulfide isomerases (PDIs) were increased in barley and B. napus, but decreased in Agrostis stolonifera, Q. robur, and poplar. Additionally, ER-luminal binding protein (BiP), trigger factor-like protein (TIG), most heat shock proteins (HSPs), and other molecular chaperones (i.e., calnexin, endoplasmin) were increased, but T-complex protein and HSP70-HSP90 organizing protein were decreased in drought-treated leaves (Supplementary Table S1). These proteins function to maintain normal protein folding, repairing, and renaturation of the stress-damaged proteins. Among them, HSPs function in protein folding for drought tolerance, which has been widely discussed [114,115,116]. HSP genes have been transferred into Arabidopsis and yeast to improve their drought tolerance [117,118]. These data suggest that maintaining correct protein folding is important for leaves to cope with drought stress [35].
Besides, protein degradation is important to remove abnormal or damaged proteins and to control the levels of certain regulatory proteins during drought stress. Proteomic studies also revealed that some protein degradation-related proteins increased in response to drought stress, including proteasomes, proteases, and peptidases (Figure 2F). Previous studies have reported that some components in the protein degradation pathway, such as ubiquitin/26S proteasomes, small ubiquitin-like modifier (E3 SUMO) ligase, and proteases/peptidases were involved in plant drought tolerance [119,120,121,122,123]. For the ubiquitin/26S proteasome system, 7 out of 11 20S proteasomes (the core regulatory particle of 26S proteasome) were increased in leaves of P. vulgaris [49], Hordeum vulgare [26], B. napus [47], and Medicago sativa [50], respectively. Importantly, the phosphorylation level of E3 ubiquitin ligase, which is one of the key enzymes involved in ubiquitination, exhibited significantly increased values in leaves under drought stress [40,44]. Many studies have shown that E3 ubiquitin ligases were positively related to plant drought tolerance [124,125,126]. These findings indicate that the enhancement of the ubiquitin/26S proteasome system is important for plants to cope with drought. In addition, most proteases were drought-increased in plants, such as ATP-dependent Clp protease in H. vulgare, cysteine proteinase in P. vulgaris, zinc metalloprotease in B. napus, and aspartic proteinase in Z. mays. Consistently, some peptidases, such as serine carboxypeptidase in G. herbaceum and oligopeptidase in O. sativa, were increased in drought-stressed leaves. Among them, aminopeptidases (APs), catalyzing the hydrolysis of amino acids from the N-terminus of proteins, were generally increased in drought-tolerant plant species, and decreased in drought-sensitive plant species or cultivars. The essential role of APs in plant drought tolerance has been well addressed previously [123].

5. Reactive Oxygen Species (ROS) Scavenging Pathways

5.1. Superoxide Dismutase (SOD), Catalase (CAT), and Peroxidase (POD) Pathway

Water deficit interrupts normal cellular metabolism that results in the production of ROS. Plants have evolved diverse mechanisms to keep ROS homeostasis in cells, including antioxidative enzymes (e.g., SOD and CAT) and chemical antioxidants (e.g., glutathione and ascorbate) (Figure 3A). Among them, SOD acts as the first line of defense by converting O2●− into H2O2, and CAT converts H2O2 into H2O and O2. They are both involved in plants’ drought tolerance. For example, the transgenic alfalfa expressing MnSOD gene from Nicotiana plumbaginifolia improved its survival and vigor after exposure to water deficit [127]. Overexpression of a cytosolic copper-zinc SOD from the mangrove plant Avicennia in rice can enhance its drought tolerance [128]. In addition, the increase of CAT activity is positively related to the drought degree [129], although the functional analysis of CAT gene on drought tolerance in transgenic plants is still scarce. The diverse abundances of SODs indicated that their functional state is dependent on subcellular location (e.g., plastid, peroxisome, or cytosol), drought conditions, and the plant’s drought adaptation ability. For example, the cytosolic Cu-Zn SODs were increased in two O. sativa cultivars (i.e., drought avoidance CT9993 and drought tolerance IR62266), while the chloroplast Cu-Zn SODs were increased in CT9993, but decreased in IR62266 [30]. Additionally, in a drought-sensitive cultivar of Malus domestica, the abundance of Cu-Zn SOD was decreased, but the FeSOD was increased [64].
POD catalyzes the reduction of H2O2 using various electron donors such as phenolic compounds, lignin precursors, auxin, and secondary metabolites. The drought-induced activities of POD were observed in leaves from Arabidopsis [130], Ramonda serbica [131], and M. sativa [132]. In proteomic studies, the abundances of PODs were increased in leaves of Quercus ilex [62] and a drought-tolerant O. sativa cultivar [32]. Moreover, the expression of POD genes was also increased in Tamarix hispida under drought stress [133]. These results indicate that PODs are critical for ROS scavenging in drought-stressed plants.

5.2. Ascorbate-Glutathione (AsA-GSH) Pathway

The AsA-GSH pathway is another key antioxidant pathway in response to drought [134,135]. During this process, the ascorbate peroxidase (APX) reduces H2O2 to H2O using ascorbate (AsA) as an electron donor, then the oxidized AsA is restored by monodehydroascorbate reductase (MDHAR), dehydroascorbate reductase (DHAR), and glutathione reductase (GR) [136] (Figure 3A). Proteomic studies revealed that most proteins in this pathway were increased under drought stress conditions. For example, the APXs localized in thylakoids and mitochondria were drought-increased in many herbaceous plants (e.g., Cynodon dactylon [19], G. max [17], T. aestivum [35,37,38], O. sativa [43,44,46], H. vulgare [37], and B. napus [47]). The function of APX under drought stress has been well addressed [129]. APX1-deficient mutant (apx1) of A. thaliana was significantly sensitive to drought stress [137]. Transgenic tobacco plants overexpressing cytosolic APX alleviated the damage from water stress. In addition, rice Osapx2 mutants had lower APX activity and were sensitive to drought, whereas overexpression of Osapx2 in rice enhanced its stress tolerance [138].
Another AsA-GSH pathway-related enzyme, GR, catalyzes the reduction of glutathione disulfide (GSSG) to the sulfhydryl form GSH. Consistently with the drought-increased abundance of GR, the drought-induced GR genes were also found in cowpea and P. vulgaris under drought stress [139,140]. Overexpression of B. rapa BrGR in E. coli showed an increase of GR activity and tolerance to H2O2 [141]. Thus, the positive function of GR in plant drought tolerance can be presumed [142]. Besides, two enzymes that can maintain ascorbate in its reduced state, dehydroascorbate reductase (DHAR) and monodehydroascorbate reductase (MDHAR), were found to be increased in some plants (i.e., C. dactylon and O. sativa) under drought stress. The expression of DHAR is generally associated with plant drought tolerance [143,144]. Overexpression of DHAR has been shown to increase drought tolerance or biomass in transgenic potato [145], rice [146], and tobacco [147]. These imply that DHAR can be a very promising target to improve plant drought stress tolerance.
In addition, proteomic studies revealed that some proteins were involved in glutathione-mediated ROS scavenging: glyoxalase (GLO), phospholipid hydroperoxide glutathione peroxidase (PHGPX), glutamate-cysteine ligase (GCL), glutaredoxin (Grx), and monothiol Grx. GLO catalyzes the detoxification of methylglyoxyl, which is involved in the indirect scavenging of ROS. Among them, the abundances of GLOs were not always increased in plants under drought stress. Their response to drought was dependent on the plant genotype and duration of drought. Besides, GCL is the first enzyme in the GSH biosynthetic pathway which was increased in drought-stressed B. napus [47]. Similar profiles were also observed in monothiol Grx in B. napus [47], Z. mays [41], as well as PHGPX in Citrullus lanatus [48]. All these suggest that the GSH-mediated antioxidative defense pathway is increased in leaves as a drought adaptation mechanism.

5.3. Peroxiredoxin/Thioredoxin (Prx/Trx) and Glutathione Peroxidase/Glutathione S-Transferase (GPX/GST) Pathway

The Prx/Trx pathway plays a central role in detoxification of H2O2. In this process, Prx catalyzes the reduction of H2O2, and then uses Trx to restore its catalytic activity (Figure 3A). The abundances of Prx and Trx were increased in response to drought, indicating the enhanced detoxification in drought-stressed leaves. Besides, a Trx-linked enzyme, methionine sulfoxide reductase (MSR), was also increased in some plants [57]. MSR is involved in the enzymatic conversion of methionine sulfoxide to methionine, which is involved in the protection of cells and tissues from H2O2-induced stress [148]. It has been reported that various oxidative stresses, including drought stress, can enhance the expression of numerous Msr genes [149]. This implies that MSR plays an important role in response to drought stress.
Besides, the GPX/GST pathway is enhanced in some plants under water deficit conditions. GPX catalyzes the reduction of H2O2 using Trx [150], and GST catalyzes conjugation reactions between GSH and a number of xenobiotics, playing a crucial role in the degradation of toxic substances. Previous studies have reported that overexpression of the GST gene enhanced drought tolerance in tobacco [151,152] and Arabidopsis [153]. Similarly, overexpression of P. glaucum GPX in rice resulted in an increased tolerance to drought stress [154]. In proteomic studies, the GPXs were drought-increased in Boea hygrometrica [21], E. elongatum [20], and B. napus [47], while drought-increased GSTs were widely found in the plant species discussed in this review. The increases of GPX and GST reinforced the functional evidence for their potential detoxification role in drought tolerance.

6. Pathogenesis Related Proteins

In proteomics results, several pathogenesis-related proteins were increased in response to drought stress, such as chitinase, disease resistance protein (DRP), polyphenol oxidase (PPO), oryzacystain, pathogen defense-related protein 10 (PR10), and disease resistance gene analog PIC15. Chitinase functions in the determent of herbivory and pathogen attack by acting on insect exoskeletons and fungal cell walls [155]. Five basic chitinase species and an acidic chitinase were increased in two tree species, i.e., Eucalyptus sp. [55] and Musa paradisiaca [42], respectively. The induced transcript of basic chitinase was also observed in Eucalyptus sp. [55]. Besides, PPO catalyzes the oxygen-dependent oxidation of phenols to quinones during plant defense against pests and pathogens. It was reported that the abundances and activities of PPOs were also increased in leaves of B. hygrometrica under desiccation [21], indicating its substantial role in the adaptation to severe water deficit. Moreover, an oryzacystain was increased in drought-stressed O. sativa leaves. Oryzacystain is a cysteine proteinase inhibitor in the phytocystatin family of proteinase inhibitors, which plays an important role in plant defense against pathogen attacks and oxidative stress [156]. The tobacco overexpressing the oryzacystain gene displayed an increase of drought tolerance by improving total SOD and guaiacol POD activities [157]. In addition, the drought-induced PR10 and disease resistance gene analog PIC15 were also found in leaves (Supplementary Table S1). It was found that overexpression of PR10 can improve drought tolerance in tobacco [158]. The response of these pathogenesis-related proteins to drought suggests that there is a cross-talk between biotic and abiotic stresses.

7. Osmotic Regulation

Osmotic regulation is crucial for plant drought resistance. Several important osmotic homeostasis-related proteins, such as late embryogenesis abundant (LEA) protein, dehydrin (DHN), and betaine aldehyde dehydrogenase (BADH), were accumulated in leaves under drought stress. Among them, LEA proteins are highly hydrophilic proteins which function in plant abiotic stress as cellular protectants to stabilize cellular components in response to water loss [159]. Similarly, in physiological studies [160], DHNs (also known as group 2 LEA proteins) were widely drought-accumulated in several plant species, including Z. mays [41], C. dactylon [19], T. aestivum [35], and B. napus [47]. Higher hydrophilicity and thermostability of DHNs suggest that they can stabilize the protein structure through detergent- and chaperone-like properties [161]. Besides, the DHN in Z. mays exhibited a significantly increased phosphorylation level under drought stress [43]. The phosphorylation of LEA2 may facilitate its binding to calcium. LEA2 acts as calcium buffer, and has calcium-dependent chaperone-like activity, which is similar to that of calreticulin and calnexin [162]. Besides, group 3 LEA proteins were also increased in Z. mays [41] and B. napus [47] under certain drought conditions. It has been reported that transformation of the LEA gene into a number of plant species can confer tolerance to drought stress. For example, the transgenic calli overexpressing sweet potato LEA14 (IbLEA14) enhanced the plant tolerance to drought stress, whereas RNA interference (RNAi) calli exhibited increased stress sensitivity [163]. All these indicate that LEA can be taken as a candidate gene for improving plant drought tolerance. Another osmotic regulation-related enzyme, BADH, showed drought accumulation in leaves of H. vulgare [27] and Hordeum spontaneum [25]. BADH converts betaine aldehyde to glycine betaine, which mediates osmotic homeostasis, and has positive effects on plant adaptation to drought stress [164]. RNAi-directed downregulation of OsBADH1 resulted in a decrease of stress tolerance and an increase of oxidative stress [165], implying its important role in plant drought tolerance.

8. Modulation of Cell Structure and Cell Cycle

8.1. Cytoskeleton and Cell Cycle

In leaves, water deficit rapidly decreases cell division rate [166,167]. Proteomic studies revealed that the abundances and phosphorylation levels of several proteins involved in cell division (i.e., cell division cycle protein, division protein ftsZ1, and cyclin A2) were decreased under water deficit [22,28,44,47] (Figure 3C). Because cytokinesis requires new cytoskeleton and cell wall components, it is predictable that the cell wall and cytoskeleton-related proteins would change in response to drought. The cytoskeleton (i.e., microtubules and actin filaments) is a highly dynamic component, which is crucial for cell division, movement, morphogenesis, and signal transduction [168]. In proteomic studies, several cytoskeleton proteins, such as actin, kinesin motor protein, tubulin, profilin, actin depolymerizing factor, and fibrillin were decreased in response to drought stress (Figure 3B). The actin genes were also reduced in drought-treated H. vulgare leaves [168]. The decrease of the cytoskeleton and cell cycle-related proteins implies that cell growth is suppressed during drought stress.
In response to drought, the translationally-controlled tumor protein homolog (TCTP) was significantly drought-increased in H. vulgare [26], T. aestivum [37], and B. napus [47], which would facilitate plant adaptation to drought stress. TCTP is a Ca2+-binding protein with important functions in some cellular processes (e.g., protection against stress and apoptosis, cell growth, cell cycle progression, and microtubule organization) [169]. The protein abundance and gene expression of TCTP were increased in P. indica-colonized plants under drought stress. Additionally, PEG treatment for 48 h also increased the abundance of TCTP in leaves of T. aestivum [37]. Overexpression of Arabidopsis TCTP enhanced drought tolerance by rapid ABA-mediated stomata closure [170], while a TCTP knockdown mutant of Arabidopsis showed severe growth defects [171]. These results imply the importance of TCTP for plant drought tolerance.

8.2. Cell Wall Modulation

Water loss in plant tissues controls turgor pressure and directly affects the extensibility of the cell wall. Some drought-responsive enzymes involved in cell wall polysaccharide synthesis/hydrolysis, lignin biosynthesis, and cell wall loosening in leaves have been identified in proteomic studies (Figure 3D). Two enzymes involved in cell wall polysaccharide synthesis, reversibly glycosylated polypeptide and pectinacetylesterase, were drought-increased in M. sativa [50] and B. napus [47], respectively. Two enzyme inhibitors involved in polysaccharide hydrolysis inhibition, the xylanase inhibitor and polygalacturonase inhibitor, were decreased in drought-sensitive Z. mays cultivars, but increased in leaves of a tolerant cultivar [41]. Their changes in response to drought stress were in contrast to that of glucan exohydrolase, which is an enzyme involved in polysaccharide hydrolysis. These results indicate that the cell wall synthesis is enhanced in a drought-tolerant maize cultivar under drought stress, which may be associated with drought adaptation. Besides, three lignin biosynthesis related proteins, phenylalanine ammonia-lyase (PAL), caffeic acid 3-O-methyltransferase, and caffeoyl-CoA O-methyl-transferase, were generally increased under drought stress (Figure 3D). PAL catalyzes the transformation of phenylalanine to cinnamylate in the first step of lignin biosynthesis [172]. The activity of PAL was also obviously increased in the leaves of Trifolium repens under the early stages of drought stress (0–14 days), but decreased gradually as the period of stress was extended [173]. In addition, two drought-increased cell wall structural proteins (i.e., glycine-rich protein and fasciclin-like arabinogalactan protein) were found in the leaves of B. napus [47], which would enhance cell wall synthesis in response to drought stress. Interestingly, a significantly increased abundance and phosphorylation level of sucrose synthase was found in Z. mays [27,32,41,44], which would facilitate the synthesis of cell wall components by providing UDP-glucose directly to the cellulose synthases and/or callose synthases [174]. The drought-increased cell wall synthesis would enhance the mechanical strength for minimizing water loss and cell dehydration, which is crucial for plants to resist and recover from drought.
Besides, cell wall loosening is also important for growth adaptation in expanding leaves [166]. Two cell wall loosening/expansion-related enzymes, polygalacturonase/pectin depolymerase (PG) in O. sativa [29] and xyloglucan endotransglycosylase (XTH) in Z. mays [41], were increased under drought stress. PG can degrade pectin, while XTH can cleave and reform the bonds between xyloglucan chains to regulate cell wall rigidity. Previous studies have found drought-induced XTH genes upregulated in Arabidopsis and rice [175,176], which implies that XTH might serve as a stress marker gene in leaves. Taken together, the cell wall modulation, which results in either growth arrest in drought-sensitive cultivars or a continuation of growth with a reduced rate [177], would contribute to drought adaption by cell size adjustment for cell turgor maintenance [17].

9. Membrane Trafficking

Plant membrane transport system plays a significant role in response to water scarcity. In proteomic studies, several membrane trafficking proteins localized in mitochondrion, plasma, and vacuole were changed in response to drought. Two mitochondrial carrier proteins (i.e., dicarboxylate/tricarboxylate carrier (DTC) and 2-oxoglutarate/malate carrier protein (OMC)) were decreased in T. aestivum [35] and G. max [17], respectively. DTC and OMC can catalyze the transport of various metabolites (e.g., dicarboxylates, tricarboxylates, amino acids, and keto acids) across the inner mitochondrial membrane, and play an important role in several metabolic processes, such as the gluconeogenesis, nitrogen metabolism, as well as biotic stress [178]. However, three voltage-dependent anion channel proteins (VDAC) in T. aestivum [35] and two mitochondrial outer membrane porin 1-like proteins in B. napus [47] were significantly increased under drought stress. The VDAC localized in the outer membrane of mitochondria can regulate Ca2+ fluxes, ATP/ADP exchange, and metabolites. The abundance changes of these mitochondrion transport-related proteins under drought stress indicated that the ion/metabolite exchange between mitochondria and cytosol was modulated in leaves to cope with the stress. Among them, the function of VDAC has been proved to be involved in plant drought tolerance [179,180]. The transcriptional level of VDAC was induced in Brassica rapa by drought stress [179], and the overexpression of the AtVDAC2 gene can confer drought resistance in Arabidopsis [180].
Remorin is a plant-specific plasma membrane protein that plays important role in plant-microbe interaction and signal transduction [181]. Some drought-increased remorins were found in Z. mays [41], O. sativa [44], and tolerant cultivars of T. aestivum [35]. Overexpressing heterologous remorin in Arabidopsis enhanced the tolerance to dehydration and salinity at the germination and seedling stages [182]. Besides, aquaporin (AQP), which transports water and other small molecules, is crucial for plants to combat drought stress. PEG treatment upregulated the expression of a PIP2 subgroup gene, AQP (TaAQP7), and the overexpression of TaAQP7 increased drought tolerance in tobacco [183]. Similarly, a proteomic study also found an increased abundance of AQP in Z. mays under 12.5% soil water stress for 6 days [41]. Additionally, drought-induced abundance of lipid transfer protein (LTP) was observed in leaves of C. albidus [53]. LTP has been proved to be involved in abiotic stress [184]. In Arabidopsis, the loss-of-function mutant of LTP3 was sensitive to drought stress, whereas overexpressing plants appeared drought tolerant [185]. Moreover, the OsDIL-overexpressing transgenic O. sativa plants were more tolerant to drought stress during vegetative and reproductive development [186]. All these demonstrated that LTP was involved in plant tolerance to drought stress, and is probably a candidate gene for genetic improvement of crop yield in adaption to drought stress.
In addition, vacuolar H+-pyrophosphatase (V-PPase), vacuolar-ATPase (V-ATPase), and ABC transporter ATPase showed dynamic changes in response to drought. V-ATPase and V-PPase are two tonoplast proton pumps for translocating H+ into the vacuoles to generate a gradient of H+, which provide a driving force for the accumulation of ions and other solutes in the vacuole. Their functions in plant abiotic stress tolerance have been widely discussed [187,188]. For example, an M. domestica vacuolar H+-ATPase (VHA) encoding gene (MdVHA-A) was induced in shoots under PEG treatment, and overexpression of MdVHA-A in transgenic tobacco seedlings improved drought tolerance [189]. In Z. mays, the abundances of V-ATPase and V-PPase were increased in a drought tolerant cultivar, but decreased in an intolerant one [41]. The differential expression patterns of ATPase and V-PPase further indicate their important functions in plants’ response to osmotic stress. Taken together, all the changes of membrane trafficking-related proteins in mitochondrion, plasma, and vacuole highlight that ion transport and membrane trafficking are crucial in leaves in order to cope with water deficit.

10. Photosynthesis and Photorespiration

It is well known that photosynthetic inhibition is one of the primary detrimental effects of water stress due to stomatal closure [190,191]. Thus, it is predictable that the universal decrease trends of photosynthesis-related proteins would be found in drought-stressed leaves (Figure 4A,B). Plants have developed many strategies to cope with drought stress, one important aspect is the recovery of photosynthesis. The drought-increased proteins involved in photoreaction and Calvin cycle were observed in leaves. For example, light-harvesting chlorophyll a/b-binding proteins (LHCB) were increased in tolerant genotypes of Z. mays [41], and M. domestica [64], but decreased or stable in sensitive genotypes. The LHCBs have been predicted to be involved in ABA signaling partially by modulating ROS homeostasis [192], and may be taken as interesting targets for crop breeding [4]. Besides, the increased abundance of sedoheptulose-1,7-bisphosphatase (SBPase) and carbonic anhydrase (CA) were found in drought-stressed P. pratensis [23] and M. domestica [64]. SBPase has a key role in regulating the photosynthetic Calvin cycle. CA catalyzes the reversible hydration of CO2, and has a relevant role in CO2 exchange by influencing the internal conductance [193].
In addition, the abundances of proteins involved in photorespiration were also decreased in some plant species, reflecting the drought-inhibited photorespiration process (Figure 4C). However, significant increases of four key enzymes (e.g., glycolate oxidase, glycine dehydrogenase, serine glyoxylate aminotransferase, and serine transhydroxymethyl transferase) were found in a high water-use efficiency M. domestica variety under moderate drought conditions [64]. Additionally, the protein abundances of aminomethyl transferase (AMT) and glycine dehydrogenase were decreased in a drought-sensitive cultivar of C. dactylon, but increased or stable in a tolerant cultivar [19]. Similarly, the downregulation of the gene encoding AMT was reported in dehydration-sensitive H. vulgare, but not in a dehydration-tolerant cultivar [194]. It is well known that photorespiration can protect photosynthesis from photoinhibition and prevent ROS accumulation in green tissues [195]. The proteomic findings provided further evidences that the regulation of photorespiration was important for plant drought tolerance [64].

11. Carbohydrate and Energy Metabolism

Proteomic studies have revealed that nearly 20% of total drought-responsive proteins were involved in carbohydrate and energy metabolism (e.g., glycolysis, tricarboxylic acid (TCA) cycle, electron transport chain, and ATP synthesis) in leaves to cope with drought stress (Figure 4E–G). For example, phosphoglucomutases (PGluMs) involved in glycolysis/gluconeogenesis were decreased in leaves of S. stapfianus [24] and O. sativa (stop irrigation for 35 days) [32], but increased in leaves of C. lanatus [48], O. sativa (under 30% relative humidity treatment for 144, 168 and 192 h) [29], and M. paradisiaca [46] in response to drought. Fructose-bisphosphate aldolase (FBPA), another glycolysis-related protein, was decreased in many aforementioned plant species, but also found to be drought-increased in other plant species such as O. sativa [29,32], S. stapfianus [24], and M. paradisiaca [46]. Besides, FBPA abundance was decreased in drought-sensitive cultivars of M. domestica [64] and P. pratensis [23], but was increased in tolerant cultivars. Similarly, aconitate hydratases involved in the TCA cycle were increased in leaves of tolerant cultivars of Z. mays and P. pratensis [23], whereas they were decreased in sensitive cultivars of Z. mays and P. pratensis [23]. The inhibition would be a mechanism for accumulating sugars as an osmolyte or energy source for recovery [62], while the increase of glycolysis and TCA may act as a strategy for providing energy during the activation of stress defenses, especially when the photosynthesis was inhibited. These apparently contradictory results could also be related to the different degrees and duration of water shortage, as well as to the tolerance ability of the plant species [62]. Among these proteins, cytosolic FBPA and glyceraldehyde-3-phosphate dehydrogenase (GAPDH) have a significantly positive correlation with drought tolerance [196], and overexpressing GAPDH in potato displayed an improvement of drought tolerance [197]. In addition, the reverse phosphorylation of some carbohydrate metabolism-related proteins (e.g., phosphoenolpyruvate carboxykinase [44], FBPA [44], and NAD-malate dehydrogenase (MDH) [34]) (Supplementary Table S2) further demonstrated that the carbon metabolism was modulated in plants in response to water stress. Although many drought-responsive carbohydrate metabolism-related proteins were revealed by proteomic studies, only a few of them have been functionally characterized to evaluate their role in drought stress tolerance.
Consistent with carbohydrate metabolism changes, mitochondrial electron transport chain and ATP synthesis-related proteins were changed in plants under drought stress. They were comprised of ATP synthase subunits (e.g., ATPase F0, F1, α, β, γ, δ, and ε subunits) and electron transport chain-related proteins (NADH dehydrogenase [47,86], quinone oxidoreductase [35,49,62], and cytochrome c oxidase [18,39,47]). Their changes imply that diverse changes of energy metabolism take place in leaves in response to water deficit. In some cases, plants may have the ability to enhance energy production to maintain major physiological activity and inhibit stress damage.

12. Nitrogen Assimilation and Amino Acid Metabolism

Nitrogen assimilation is affected by abiotic stress in plants. In this process, exogenous absorbed nitrate is transformed to ammonium by nitrate reductase (NR) and nitrite reductase (NIR), and then assimilated by glutamine synthetase (GS) and glutamate synthase (GOGAT) into amino acids. Proteomic studies revealed that several enzymes in nitrogen assimilation were drought-decreased in leaves (Figure 4H). Long-term (3 weeks) drought stress decreased the abundance of NIR in H. vulgare [25], but short-term (2–5 days) drought stress increased NIR in O. sativa [29]. Besides, drought significantly decreased the protein abundances of GS1, GS2, and GOGAT in many plant species (Supplementary Table S1), and the activities of NR and GS were reduced in Leymus chinensis [198], T. aestivum [199], and tomato [200] under water deficit. Similarly, the decrease at transcriptional level of Brassica juncea GS1.1 and other key genes involved in nitrogen assimilation (e.g., BjGDH1, BjGDH2, and BjASN2) were also well addressed [201]. The reduction of NR, GS, and GOGAT indicates that the nitrogen assimilation is significantly inhibited by drought stress, which is supposed to be the main reason for the reduction of crop yield under drought stress.
Following the GS/GOGAT cycle, the amino group was transferred into other amino acids by aminotransferases (Figure 4H). Two aminotransferases (e.g., aspartate aminotransferase (AST) and alanine aminotransferase (ALAT)) were altered under drought stress. Among them, AST was drought-decreased in P. pratensis [23], and ALAT was decreased in P. pratensis [23] and Z. mays [41]. AST catalyzes the amino transfer from glutamate to oxaloacetate to form aspartate, which is part of ammonium assimilation. ALAT catalyzes the interconversion of l-glutamate and l-alanine. The abundance decline of these enzymes indicates that the amino acid metabolism and the synthesis of other metabolites derived from amino acids are inhibited under drought stress.
Interestingly, several key enzymes involved in S-adenosyl-l-methionine (SAM) cycle were generally increased in leaves in response to drought stress, including drought-increased 5-methyltetrahydropteroyltriglutamate-homocysteine methyltransferase (MetE), S-adenosyl-l-homocysteine hydrolase (SAHase), S-adenosylmethionine synthase (SAMS), and methionine synthase (MS) (Figure 4H). Similarly, MS in Z. mays [41], Sorghum bicolor [202], a drought-tolerant C. dactylon cultivar, as well as SAMS in Populus [59,86], Q. ilex [63], and T. aestivum [35] were increased under drought conditions. The metabolites in the SAM cycle were known to play major roles in methylation of DNA, proteins, and other metabolites, being involved in control of gene expression, cell wall metabolism, as well as the biosynthesis of polyamine and ethylene for stress tolerance [18,203,204]. In addition, the MS gene was also strongly induced in barley leaves under drought stress [205]. Transgenic Arabidopsis lines overexpressing SAMS from Solanum brevidens enhanced the ability of plant drought tolerance [206]. This implies that the increases of MS and SAMS could enhance the methionine and osmoregulant metabolism for plants to cope with drought stress.

13. Fatty Acid Metabolism and Other Metabolisms

Four proteins involved in fatty acid biosynthesis (i.e., acetyl-coenzyme A carboxylase carboxyl transferase, acyl carrier protein, enoyl-acyl carrier protein reductase, and lipoxygenase 6) and three involved in fatty acid degradation (i.e., thiolase I, thiolase II, and acyl-CoA dehydrogenase) were altered under dehydration conditions (Figure 4I). Physiological and molecular investigations have revealed that the fatty acid/lipid metabolism was changed in response to drought stress [207,208,209]. For example, a greater composition of unsaturated fatty acids in membrane lipids may contribute to superior leaf dehydration tolerance in C. dactylon [209]. In the leaves of Parkinsonia aculeata, the significantly increased contents of chloroplast lipids (i.e., monogalactosyldiacylglycerol and digalactosyldiacylglycerol) and extra-plastidial lipids (i.e., phosphatidylcholine and phosphatidylethanolamine) maintained Parkinsonia plants in a state of rapid recovery of lipid synthesis during rehydration [209]. The changes in fatty acid/lipid composition may help to maintain membrane integrity and preserve cell compartmentation under water stress [210].
In addition, two flavonoid biosynthesis related proteins (i.e., chalcone isomerase (CHI) and dihydroflavonol-4-reductase) involved in secondary metabolism were also changed in response to drought (Supplementary Table S1). It has been suggested that flavonoids have antioxidant capacity to protect plants against abiotic stress [211,212], and flavonoid accumulation was induced rapidly by abiotic stresses [213]. In proteomic studies, CHI, a key enzyme in flavonoid biosynthesis, was decreased in O. sativa under drought stress [29]. Similarly, the CsCHI gene in leaves of tea (Camellia sinensis) was also downregulated under drought treatment [214]. Besides, dihydroflavonol-4-reductase, another key enzyme in the flavonoid biosynthesis pathway, was decreased in a drought sensitive Z. mays cultivar [41], but increased in a tolerant one. The abundance changes of these enzymes currently attract great attention from researchers seeking to evaluate their potential protective roles in drought stress response.

14. Concluding Remarks

By integrative analysis of proteomics results, it has been found that, in total, 2216 protein species and more than 200 phosphoproteins were drought-responsive in plant leaves. The abundance changes of these proteins provide new clues towards further understanding the complex cellular and molecular processes in plant drought tolerance. The quantitative patterns of protein in various plants give rise to diverse drought-responsive mechanisms, including drought signal transduction, ROS scavenging, osmotic regulation, specific gene expression regulation, protein synthesis and turnover, cell structure modulation, as well as carbohydrate and energy metabolism (Figure 5). Exploring these protein functions will enable a holistic understanding of plant-environment interaction. Moreover, the potential proteins screened out by proteomics will provide candidate genes/proteins for genetic improvement in plant drought tolerance. However, some proteins showed dynamic changes depending on plant species and stress intensity, which makes it difficult to give a clear interpretation of the mechanism in plant drought response. Importantly, proteomic investigations also revealed that protein PTMs are crucial for signaling and metabolic mechanisms in plant drought response. The integration of those findings from physiological, gene expression, and other large-scale “omics” will help us to establish molecular networks of drought response and tolerance.

Supplementary Materials

Supplementary materials can be found at www.mdpi.com/1422-0067/17/10/1706/s1.

Acknowledgments

This work was supported by the Natural Science Foundation of Shanghai (No. 15ZR1431300), Capacity Construction Project of Local Universities, Shanghai, (No. 14390502700), and Shanghai Municipal Agricultural Commission (No. 201509).

Author Contributions

Xiaoli Wang and Shaojun Dai designed this work, collected the data, and co-wrote the manuscript. Xiaofeng Cai, Chengxi Xu, and Quanhua Wang collected the data and helped in editing the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chaves, M.M.; Pereira, J.S.; Maroco, J.; Rodrigues, M.L.; Ricardo, C.P.P.; Osorio, M.L.; Carvalho, I.; Faria, T.; Pinheiro, C. How plants cope with water stress in the field. Photosynthesis and growth. Ann. Bot. 2002, 89, 907–916. [Google Scholar] [PubMed]
  2. Dinakar, C.; Bartels, D. Desiccation tolerance in resurrection plants: New insights from transcriptome, proteome, and metabolome analysis. Front. Plant Sci. 2013, 4, 482. [Google Scholar] [CrossRef] [PubMed]
  3. Jogaiah, S.; Govind, S.R.; Tran, L.S.P. Systems biology-based approaches toward understanding drought tolerance in food crops. Crit. Rev. Biotechnol. 2013, 33, 23–39. [Google Scholar] [CrossRef] [PubMed]
  4. Abreu, I.A.; Farinha, A.P.; Negrao, S.; Goncalves, N.; Fonseca, C.; Rodrigues, M.; Batista, R.; Saibo, N.J.M.; Oliveira, M.M. Coping with abiotic stress: Proteome changes for crop improvement. J. Proteom. 2013, 93, 145–168. [Google Scholar] [CrossRef] [PubMed]
  5. Farooq, M.; Wahid, A.; Kobayashi, N.; Fujita, D.; Basra, S.M.A. Plant drought stress: Effects, mechanisms and management. Agron. Sustain. Dev. 2009, 29, 185–212. [Google Scholar] [CrossRef]
  6. Osmond, C.B.; Grace, S.C. Perspectives on photoinhibition and photorespiration in the field: Quintessential inefficiencies of the light and dark reactions of photosynthesis? J. Exp. Bot. 1995, 46, 1351–1362. [Google Scholar] [CrossRef]
  7. Zhang, Q. Strategies for developing Green Super Rice. Proc. Natl. Acad. Sci. USA 2007, 104, 16402–16409. [Google Scholar] [CrossRef] [PubMed]
  8. El-Soda, M.; Kruijer, W.; Malosetti, M.; Koornneef, M.; Aarts, M.G.M. Quantitative trait loci and candidate genes underlying genotype by environment interaction in the response of Arabidopsis thaliana to drought. Plant Cell Environ. 2015, 38, 585–599. [Google Scholar] [CrossRef] [PubMed]
  9. Yadav, R.S.; Sehgal, D.; Vadez, V. Using genetic mapping and genomics approaches in understanding and improving drought tolerance in pearl millet. J. Exp. Bot. 2011, 62, 397–408. [Google Scholar] [CrossRef] [PubMed]
  10. Alm, V.; Busso, C.S.; Ergon, A.; Rudi, H.; Larsen, A.; Humphreys, M.W.; Rognli, O.A. QTL analyses and comparative genetic mapping of frost tolerance, winter survival and drought tolerance in meadow fescue (Festuca pratensis Huds.). Theor. Appl. Genet. 2011, 123, 369–382. [Google Scholar] [CrossRef] [PubMed]
  11. Acuna-Galindo, M.A.; Mason, R.E.; Subramanian, N.K.; Hays, D.B. Meta-analysis of wheat QTL regions associated with adaptation to drought and heat stress. Crop Sci. 2015, 55, 477–492. [Google Scholar] [CrossRef]
  12. Hao, Z.; Li, X.; Liu, X.; Xie, C.; Li, M.; Zhang, D.; Zhang, S. Meta-analysis of constitutive and adaptive QTL for drought tolerance in maize. Euphytica 2010, 174, 165–177. [Google Scholar] [CrossRef]
  13. Asfaw, A.; Blair, M.W. Quantitative trait loci for rooting pattern traits of common beans grown under drought stress versus non-stress conditions. Mol. Breed. 2012, 30, 681–695. [Google Scholar] [CrossRef]
  14. Reddy, S.K.; Liu, S.; Rudd, J.C.; Xue, Q.; Payton, P.; Finlayson, S.A.; Mahan, J.; Akhunova, A.; Holalu, S.V.; Lu, N. Physiology and transcriptomics of water-deficit stress responses in wheat cultivars TAM 111 and TAM 112. J. Plant Physiol. 2014, 171, 1289–1298. [Google Scholar] [CrossRef] [PubMed]
  15. Krannich, C.T.; Maletzki, L.; Kurowsky, C.; Horn, R. Network candidate genes in breeding for drought tolerant crops. Int. J. Mol. Sci. 2015, 16, 16378–16400. [Google Scholar] [CrossRef] [PubMed]
  16. Baerenfaller, K.; Massonnet, C.; Walsh, S.; Baginsky, S.; Buehlmann, P.; Hennig, L.; Hirsch-Hoffmann, M.; Howell, K.A.; Kahlau, S.; Radziejwoski, A.; et al. Systems-based analysis of Arabidopsis leaf growth reveals adaptation to water deficit. Mol. Syst. Biol. 2012, 8, 606. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Mohammadi, P.P.; Moieni, A.; Hiraga, S.; Komatsu, S. Organ-specific proteomic analysis of drought-stressed soybean seedlings. J. Proteom. 2012, 75, 1906–1923. [Google Scholar] [CrossRef] [PubMed]
  18. Merewitz, E.B.; Gianfagna, T.; Huang, B. Protein accumulation in leaves and roots associated with improved drought tolerance in creeping bentgrass expressing an ipt gene for cytokinin synthesis. J. Exp. Bot. 2011, 62, 5311–5333. [Google Scholar] [CrossRef] [PubMed]
  19. Zhao, Y.; Du, H.; Wang, Z.; Huang, B. Identification of proteins associated with water-deficit tolerance in C-4 perennial grass species, Cynodon dactylon × Cynodon transvaalensis and Cynodon dactylon. Physiol. Plant. 2011, 141, 40–55. [Google Scholar] [CrossRef] [PubMed]
  20. Gazanchian, A.; Hajheidari, M.; Sima, N.K.; Salekdeh, G.H. Proteome response of Elymus elongatum to severe water stress and recovery. J. Exp. Bot. 2007, 58, 291–300. [Google Scholar] [CrossRef] [PubMed]
  21. Jiang, G.; Wang, Z.; Shang, H.; Yang, W.; Hu, Z.; Phillips, J.; Deng, X. Proteome analysis of leaves from the resurrection plant Boea hygrometrica in response to dehydration and rehydration. Planta 2007, 225, 1405–1420. [Google Scholar] [CrossRef] [PubMed]
  22. Xu, C.; Huang, B. Proteins and metabolites regulated by trinexapac-ethyl in relation to drought tolerance in Kentucky bluegrass. J. Plant Growth Regul. 2012, 31, 25–37. [Google Scholar] [CrossRef]
  23. Xu, C.; Huang, B. Comparative analysis of drought responsive proteins in Kentucky bluegrass cultivars contrasting in drought tolerance. Crop Sci. 2010, 50, 2543–2552. [Google Scholar] [CrossRef]
  24. Oliver, M.J.; Jain, R.; Balbuena, T.S.; Agrawal, G.; Gasulla, F.; Thelen, J.J. Proteome analysis of leaves of the desiccation-tolerant grass, Sporobolus stapfianus, in response to dehydration. Phytochemistry 2011, 72, 1273–1284. [Google Scholar] [CrossRef] [PubMed]
  25. Ashoub, A.; Baeumlisberger, M.; Neupaertl, M.; Karas, M.; Brueggemann, W. Characterization of common and distinctive adjustments of wild barley leaf proteome under drought acclimation, heat stress and their combination. Plant Mol. Biol. 2015, 87, 459–471. [Google Scholar] [CrossRef] [PubMed]
  26. Ghabooli, M.; Khatabi, B.; Ahmadi, F.S.; Sepehri, M.; Mirzaei, M.; Amirkhani, A.; Jorrin-Novo, J.V.; Salekdeh, G.H. Proteomics study reveals the molecular mechanisms underlying water stress tolerance induced by Piriformospora indica in barley. J. Proteom. 2013, 94, 289–301. [Google Scholar] [CrossRef] [PubMed]
  27. Ashoub, A.; Beckhaus, T.; Berberich, T.; Karas, M.; Bruggemann, W. Comparative analysis of barley leaf proteome as affected by drought stress. Planta 2013, 237, 771–781. [Google Scholar] [CrossRef] [PubMed]
  28. Wendelboe-Nelson, C.; Morris, P.C. Proteins linked to drought tolerance revealed by DIGE analysis of drought resistant and susceptible barley varieties. Proteomics 2012, 12, 3374–3385. [Google Scholar] [CrossRef] [PubMed]
  29. Pandey, A.; Rajamani, U.; Verma, J.; Subba, P.; Chakraborty, N.; Datta, A.; Chakraborty, S.; Chakraborty, N. Identification of extracellular matrix proteins of rice (Oryza sativa L.) Involved in dehydration-responsive network: A proteomic approach. J. Proteome Res. 2010, 9, 3443–3464. [Google Scholar] [PubMed]
  30. Salekdeh, G.H.; Siopongco, J.; Wade, L.J.; Ghareyazie, B.; Bennett, J. Proteomic analysis of rice leaves during drought stress and recovery. Proteomics 2002, 2, 1131–1145. [Google Scholar] [CrossRef]
  31. Rabello, F.R.; Villeth, G.R.C.; Rabello, A.R.; Rangel, P.H.N.; Guimaraes, C.M.; Huergo, L.F.; Souza, E.M.; Pedrosa, F.O.; Ferreira, M.E.; Mehta, A. Proteomic analysis of upland rice (Oryza sativa L.) exposed to intermittent water deficit. Protein J. 2014, 33, 221–230. [Google Scholar] [CrossRef] [PubMed]
  32. Raorane, M.L.; Pabuayon, I.M.; Varadarajan, A.R.; Mutte, S.K.; Kumar, A.; Treumann, A.; Kohli, A. Proteomic insights into the role of the large-effect QTL qDTY12.1 for rice yield under drought. Mol. Breed. 2015, 35, 1–14. [Google Scholar] [CrossRef]
  33. Ji, K.X.; Wang, Y.Y.; Sun, W.N.; Lou, Q.J.; Mei, H.W.; Shen, S.H.; Chen, H. Drought-responsive mechanisms in rice genotypes with contrasting drought tolerance during reproductive stage. J. Plant Physiol. 2012, 169, 336–344. [Google Scholar] [CrossRef] [PubMed]
  34. Ke, Y.Q.; Han, G.Q.; He, H.Q.; Li, J.X. Differential regulation of proteins and phosphoproteins in rice under drought stress. Biochem. Biophys. Res. Commun. 2009, 379, 133–138. [Google Scholar] [CrossRef] [PubMed]
  35. Ford, K.L.; Cassin, A.; Bacic, A. Quantitative proteomic analysis of wheat cultivars with differing drought stress tolerance. Front. Plant Sci. 2011, 2, 44. [Google Scholar] [CrossRef] [PubMed]
  36. Peremarti, A.; Mare, C.; Aprile, A.; Roncaglia, E.; Cattivelli, L.; Villegas, D.; Royo, C. Transcriptomic and proteomic analyses of a pale-green durum wheat mutant shows variations in photosystem components and metabolic deficiencies under drought stress. BMC Genom. 2014, 15, 125. [Google Scholar] [CrossRef] [PubMed]
  37. Zhang, H.M.; Zhang, L.S.; Lv, H.; Yu, Z.Y.; Zhang, D.P.; Zhu, W.N. Identification of changes in Triticum aestivum L. leaf proteome in response to drought stress by 2D-PAGE and MALDI-TOF/TOF mass spectrometry. Acta Physiol. Plant. 2014, 36, 1385–1398. [Google Scholar] [CrossRef]
  38. Wang, X.; Vignjevic, M.; Jiang, D.; Jacobsen, S.; Wollenweber, B. Improved tolerance to drought stress after anthesis due to priming before anthesis in wheat (Triticum aestivum L.) var. Vinjett. J. Exp. Bot. 2014, 65, 6441–6456. [Google Scholar] [CrossRef] [PubMed]
  39. Budak, H.; Akpinar, B.A.; Unver, T.; Turktas, M. Proteome changes in wild and modern wheat leaves upon drought stress by two-dimensional electrophoresis and nanoLC-ESI-MS/MS. Plant Mol. Biol. 2013, 83, 89–103. [Google Scholar] [CrossRef] [PubMed]
  40. Zhang, M.; Lv, D.W.; Ge, P.; Bian, Y.W.; Chen, G.X.; Zhu, G.R.; Li, X.H.; Yan, Y.M. Phosphoproteome analysis reveals new drought response and defense mechanisms of seedling leaves in bread wheat (Triticum aestivum L.). J. Proteom. 2014, 109, 290–308. [Google Scholar] [CrossRef] [PubMed]
  41. Benesova, M.; Hola, D.; Fischer, L.; Jedelsky, P.L.; Hnilicka, F.; Wilhelmova, N.; Rothova, O.; Kocova, M.; Prochazkova, D.; Honnerova, J.; et al. The physiology and proteomics of drought tolerance in maize: Early stomatal closure as a cause of lower tolerance to short-term dehydration? PLoS ONE 2012, 7, e38017. [Google Scholar] [CrossRef] [PubMed]
  42. Sanggon, K.; Jinseok, L.; Jungtae, K.; Youngsang, K.; Dongwon, B.; Hwanhee, B.; Beomyoung, S.; Seongbum, B.; Youngup, K.; Miok, W. Physiological and proteomic analysis of the response to drought stress in an inbred Korean maize line. Plant Omics 2015, 8, 159–168. [Google Scholar]
  43. Bonhomme, L.; Valot, B.; Tardieu, F.; Zivy, M. Phosphoproteome dynamics upon changes in plant water status reveal early events associated with rapid growth adjustment in maize leaves. Mol. Cell. Proteom. 2012, 11, 957–972. [Google Scholar] [CrossRef] [PubMed]
  44. Hu, X.L.; Wu, L.J.; Zhao, F.Y.; Zhang, D.Y.; Li, N.N.; Zhu, G.H.; Li, C.H.; Wang, W. Phosphoproteomic analysis of the response of maize leaves to drought, heat and their combination stress. Front. Plant Sci. 2015, 6, 289. [Google Scholar] [CrossRef] [PubMed]
  45. Zhou, G.; Yang, L.T.; Li, Y.R.; Zou, C.L.; Huang, L.P.; Qiu, L.H.; Huang, X.; Srivastava, M.K. Proteomic analysis of osmotic stress-responsive proteins in sugarcane leaves. Plant Mol. Biol. Rep. 2012, 30, 349–359. [Google Scholar] [CrossRef]
  46. Vanhove, A.-C.; Vermaelen, W.; Panis, B.; Swennen, R.; Carpentier, S.C. Screening the banana biodiversity for drought tolerance: Can an in vitro growth model and proteomics be used as a tool to discover tolerant varieties and understand homeostasis. Front. Plant Sci. 2012, 3, 176. [Google Scholar] [CrossRef] [PubMed]
  47. Koh, J.; Chen, G.; Yoo, M.-J.; Zhu, N.; Dufresne, D.; Erickson, J.E.; Shao, H.; Chen, S. Comparative proteomic analysis of Brassica napus in response to drought stress. J. Proteome Res. 2015, 14, 3068–3081. [Google Scholar] [CrossRef] [PubMed]
  48. Akashi, K.; Yoshida, K.; Kuwano, M.; Kajikawa, M.; Yoshimura, K.; Hoshiyasu, S.; Inagaki, N.; Yokota, A. Dynamic changes in the leaf proteome of a C-3 xerophyte, Citrullus lanatus (wild watermelon), in response to water deficit. Planta 2011, 233, 947–960. [Google Scholar] [CrossRef] [PubMed]
  49. Zadraznik, T.; Hollung, K.; Egge-Jacobsen, W.; Meglic, V.; Sustar-Vozlic, J. Differential proteomic analysis of drought stress response in leaves of common bean (Phaseolus vulgaris L.). J. Proteom. 2013, 78, 254–272. [Google Scholar] [CrossRef] [PubMed]
  50. Aranjuelo, I.; Molero, G.; Erice, G.; Avice, J.C.; Nogues, S. Plant physiology and proteomics reveals the leaf response to drought in alfalfa (Medicago sativa L.). J. Exp. Bot. 2011, 62, 111–123. [Google Scholar] [CrossRef] [PubMed]
  51. Deeba, F.; Pandey, A.K.; Ranjan, S.; Mishra, A.; Singh, R.; Sharma, Y.K.; Shirke, P.A.; Pandey, V. Physiological and proteomic responses of cotton (Gossypium herbaceum L.) to drought stress. Plant Physiol. Biochem. 2012, 53, 6–18. [Google Scholar] [CrossRef] [PubMed]
  52. Zhang, M.H.; Li, G.W.; Huang, W.; Bi, T.; Chen, G.Y.; Tang, Z.C.; Su, W.A.; Sun, W.N. Proteomic study of Carissa spinarum in response to combined heat and drought stress. Proteomics 2010, 10, 3117–3129. [Google Scholar] [CrossRef] [PubMed]
  53. Brossa, R.; Pinto-Marijuan, M.; Francisco, R.; Lopez-Carbonell, M.; Chaves, M.M.; Alegre, L. Redox proteomics and physiological responses in Cistus albidus shrubs subjected to long-term summer drought followed by recovery. Planta 2015, 241, 803–822. [Google Scholar] [CrossRef] [PubMed]
  54. Xu, G.; Li, C.Y.; Yao, Y.N. Proteomics analysis of drought stress-responsive proteins in Hippophae rhamnoides L. Plant Mol. Biol. Rep. 2009, 27, 153–161. [Google Scholar] [CrossRef]
  55. Bedon, F.; Villar, E.; Vincent, D.; Dupuy, J.W.; Lomenech, A.M.; Mabialangoma, A.; Chaumeil, P.; Barre, A.; Plomion, C.; Gion, J.M. Proteomic plasticity of two Eucalyptus genotypes under contrasted water regimes in the field. Plant Cell Environ. 2012, 35, 790–805. [Google Scholar] [CrossRef] [PubMed]
  56. Bonhomme, L.; Monclus, R.; Vincent, D.; Carpin, S.; Claverol, S.; Lomenech, A.-M.; Labas, V.; Plomion, C.; Brignolas, F.; Morabito, D. Genetic variation and drought response in two Populus × euramericana genotypes through 2-DE proteomic analysis of leaves from field and glasshouse cultivated plants. Phytochemistry 2009, 70, 988–1002. [Google Scholar] [CrossRef] [PubMed]
  57. He, C.Y.; Zhang, J.G.; Duan, A.G.; Zheng, S.X.; Sun, H.G.; Fu, L.H. Proteins responding to drought and high-temperature stress in Populus × euramericana cv. “74/76”. Trees Struct. Funct. 2008, 22, 803–813. [Google Scholar] [CrossRef]
  58. Durand, T.C.; Sergeant, K.; Renaut, J.; Planchon, S.; Hoffmann, L.; Carpin, S.; Label, P.; Morabito, D.; Hausman, J.F. Poplar under drought: Comparison of leaf and cambial proteomic responses. J. Proteom. 2011, 74, 1396–1410. [Google Scholar] [CrossRef] [PubMed]
  59. Bohler, S.; Sergeant, K.; Jolivet, Y.; Hoffmann, L.; Hausman, J.-F.; Dizengremel, P.; Renaut, J. A physiological and proteomic study of poplar leaves during ozone exposure combined with mild drought. Proteomics 2013, 13, 1737–1754. [Google Scholar] [CrossRef] [PubMed]
  60. Xiao, X.; Yang, F.; Zhang, S.; Korpelainen, H.; Li, C. Physiological and proteomic responses of two contrasting Populus cathayana populations to drought stress. Physiol. Plant. 2009, 136, 150–168. [Google Scholar] [CrossRef] [PubMed]
  61. Sergeant, K.; Spiess, N.; Renaut, J.; Wilhelm, E.; Hausman, J.F. One dry summer: A leaf proteome study on the response of oak to drought exposure. J. Proteom. 2011, 74, 1385–1395. [Google Scholar] [CrossRef] [PubMed]
  62. Echevarria-Zomeno, S.; Ariza, D.; Jorge, I.; Lenz, C.; del Campo, A.; Jorrin, J.V.; Navarro, R.M. Changes in the protein profile of Quercus ilex leaves in response to drought stress and recovery. J. Plant Physiol. 2009, 166, 233–245. [Google Scholar] [CrossRef] [PubMed]
  63. Valero-Galvan, J.; Gonzalez-Fernandez, R.; Navarro-Cerrillo, R.M.; Gil-Pelegrin, E.; Jorrin-Novo, J.V. Physiological and proteomic analyses of drought stress response in Holm oak provenances. J. Proteome Res. 2013, 12, 5110–5123. [Google Scholar] [CrossRef] [PubMed]
  64. Zhou, S.; Li, M.; Guan, Q.; Liu, F.; Zhang, S.; Chen, W.; Yin, L.; Qin, Y.; Ma, F. Physiological and proteome analysis suggest critical roles for the photosynthetic system for high water-use efficiency under drought stress in Malus. Plant Sci. 2015, 236, 44–60. [Google Scholar] [CrossRef] [PubMed]
  65. Carvalho, R.F.; Campos, M.L.; Azevedo, R.A. The role of phytochrome in stress tolerance. J. Integr. Plant Biol. 2011, 53, 920–929. [Google Scholar] [CrossRef] [PubMed]
  66. Boggs, J.Z.; Loewy, K.; Bibee, K.; Heschel, M.S. Phytochromes influence stomatal conductance plasticity in Arabidopsis thaliana. Plant Growth Regul. 2010, 60, 77–81. [Google Scholar] [CrossRef]
  67. Yadav, D.K.; Shukla, D.; Tuteja, N. Rice heterotrimeric G-protein alpha subunit (RGA1): In silico analysis of the gene and promoter and its upregulation under abiotic stress. Plant Physiol. Biochem. 2013, 63, 262–271. [Google Scholar] [CrossRef] [PubMed]
  68. Gao, Y.; Li, T.; Liu, Y.; Ren, C.; Zhao, Y.; Wang, M. Isolation and characterization of gene encoding G protein α subunit protein responsive to plant hormones and abiotic stresses in Brassica napus. Mol. Biol. Rep. 2010, 37, 3957–3965. [Google Scholar] [CrossRef] [PubMed]
  69. Xu, D.-B.; Chen, M.; Ma, Y.N.; Xu, Z.S.; Li, L.C.; Chen, Y.F.; Ma, Y.Z. A G-protein β subunit, AGB1, negatively regulates the ABA response and drought tolerance by down-regulating AtMPK6-related pathway in Arabidopsis. PLoS ONE 2015, 10, e0116385. [Google Scholar] [CrossRef] [PubMed]
  70. Zhang, L.; Hu, G.; Cheng, Y.; Huang, J. Heterotrimeric G protein α, and β subunits antagonistically modulate stomatal density in Arabidopsis thaliana. Dev. Biol. 2008, 324, 68–75. [Google Scholar] [CrossRef] [PubMed]
  71. Ma, Q.-H. Small GTP-binding Proteins and their Functions in Plants. J. Plant Growth Regul. 2007, 26, 369–388. [Google Scholar] [CrossRef]
  72. Choudhary, M.; Jayanand Padaria, J.C. Transcriptional profiling in pearl millet (Pennisetum glaucum L.R. Br.) for identification of differentially expressed drought responsive genes. Physiol. Mol. Biol. Plants 2015, 21, 187–196. [Google Scholar] [CrossRef] [PubMed]
  73. Song, L.; Li, R.; Xiang, X.; Wang, J.; Qiao, L.; Song, X.; Pei, Y.; Sui, J. Overexpression of stress-inducible small GTP-binding protein AhRab7 (AhRabG3f) in peanut (Arachis hypogaea L.) enhances abiotic stress tolerance. J. Food Agric. Environ. 2012, 10, 888–894. [Google Scholar]
  74. Liu, Z.; Luo, C.; Li, L.; Doug, L.; Can, V.; Wei, P.; He, X. Isolation, characterization and expression analysis of the GDP dissociation inhibitor protein gene MiRab-GDI from Mangifera indica L. Sci. Hortic. 2015, 185, 14–21. [Google Scholar] [CrossRef]
  75. Huang, G.T.; Ma, S.L.; Bai, L.P.; Zhang, L.; Ma, H.; Jia, P.; Liu, J.; Zhong, M.; Guo, Z.F. Signal transduction during cold, salt, and drought stresses in plants. Mol. Biol. Rep. 2012, 39, 969–987. [Google Scholar] [CrossRef] [PubMed]
  76. Garg, G.; Yadav, S.; Ruchi Yadav, G. Key roles of calreticulin and calnexin proteins in plant perception under stress conditions: A Review. Adv. Life Sci. 2015, 5, 18–26. [Google Scholar]
  77. Islam, M.; Begum, M.C.; Kabir, A.H.; Alam, M.F. Molecular and biochemical mechanisms associated with differential responses to drought tolerance in wheat (Triticum aestivum L.). J. Plant Interact. 2015, 10, 195–201. [Google Scholar] [CrossRef]
  78. Jia, X.Y.; Xu, C.Y.; Jing, R.L.; Li, R.Z.; Mao, X.G.; Wang, J.P.; Chang, X.P. Molecular cloning and characterization of wheat calreticulin (CRT) gene involved in drought-stressed responses. J. Exp. Bot. 2008, 59, 739–751. [Google Scholar] [CrossRef] [PubMed]
  79. Jiang, S.; Zhang, D.; Wang, L.; Pan, J.; Liu, Y.; Kong, X.; Zhou, Y.; Li, D. A maize calcium-dependent protein kinase gene, ZmCPK4, positively regulated abscisic acid signaling and enhanced drought stress tolerance in transgenic Arabidopsis. Plant Physiol. Biochem. 2013, 71, 112–120. [Google Scholar] [CrossRef] [PubMed]
  80. Zou, J.J.; Wei, F.J.; Wang, C.; Wu, J.J.; Ratnasekera, D.; Liu, W.X.; Wu, W.H. Arabidopsis calcium-dependent protein kinase CPK10 functions in abscisic acid- and Ca2+-mediated stomatal regulation in response to drought stress. Plant Physiol. 2010, 154, 1232–1243. [Google Scholar] [CrossRef] [PubMed]
  81. Wei, S.; Hu, W.; Deng, X.; Zhang, Y.; Liu, X.; Zhao, X.; Luo, Q.; Jin, Z.; Li, Y.; Zhou, S.; et al. A rice calcium-dependent protein kinase OsCPK9 positively regulates drought stress tolerance and spikelet fertility. BMC Plant Biol. 2014, 14, 133. [Google Scholar] [CrossRef] [PubMed]
  82. You, J.; Chan, Z. ROS regulation during abiotic stress responses in crop plants. Front. Plant Sci. 2015, 6, 1092. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, L.X.; Hu, X.L.; Song, J.; Zong, X.j.; Li, D.P.; Li, D.Q. Over-expression of a Zea mays L. protein phosphatase 2C gene (ZmPP2C) in Arabidopsis thaliana decreases tolerance to salt and drought. J. Plant Physiol. 2009, 166, 531–542. [Google Scholar] [CrossRef] [PubMed]
  84. Gosti, F.; Beaudoin, N.; Serizet, C.; Webb, A.A.; Vartanian, N.; Giraudat, J. ABI1 protein phosphatase 2C is a negative regulator of abscisic acid signaling. Plant Cell 1999, 11, 1897–1910. [Google Scholar] [CrossRef] [PubMed]
  85. Li, F.H.; Fu, F.L.; Sha, L.N.; He, L.; Li, W.C. Differential expression of serine/threonine protein phosphatase type-2C under drought stress in maize. Plant Mol. Biol. Rep. 2009, 27, 29–37. [Google Scholar] [CrossRef]
  86. Plomion, C.; Lalanne, C.; Claverol, S.; Meddour, H.; Kohler, A.; Bogeat-Triboulot, M.-B.; Barre, A.; Le Provost, G.; Dumazet, H.; Jacob, D.; et al. Mapping the proteome of poplar and application to the discovery of drought-stress responsive proteins. Proteomics 2006, 6, 6509–6527. [Google Scholar] [CrossRef] [PubMed]
  87. Guo, B.Z.; Xu, G.; Cao, Y.G.; Holbrook, C.C.; Lynch, R.E. Identification and characterization of phospholipase D and its association with drought susceptibilities in peanut (Arachis hypogaea). Planta 2006, 223, 512–520. [Google Scholar] [CrossRef] [PubMed]
  88. El Maarouf, H.; Zuily-Fodil, Y.; Gareil, M.; d'Arcy-Lameta, A.; Pham-Thi, A.T. Enzymatic activity and gene expression under water stress of phospholipase D in two cultivars of Vigna unguiculata L. Walp. differing in drought tolerance. Plant Mol. Biol. 1999, 39, 1257–1265. [Google Scholar] [CrossRef] [PubMed]
  89. Wang, J.; Ding, B.; Guo, Y.; Li, M.; Chen, S.; Huang, G.; Xie, X. Overexpression of a wheat phospholipase D gene, TaPLDα, enhances tolerance to drought and osmotic stress in Arabidopsis thaliana. Planta 2014, 240, 103–115. [Google Scholar] [CrossRef] [PubMed]
  90. Zhang, T.; Song, Y.; Liu, Y.; Guo, X.; Zhu, C.; Wen, F. Overexpression of phospholipase gene enhances drought and salt tolerance of Populus tomentosa. Chin. Sci. Bull. 2008, 53, 3656–3665. [Google Scholar] [CrossRef]
  91. Hong, Y.; Zheng, S.; Wang, X. Dual functions of phospholipase Dα1 in plant response to drought. Mol. Plant 2008, 1, 262–269. [Google Scholar] [CrossRef] [PubMed]
  92. Fu, H.; Subramanian, R.R.; Masters, S.C. 14-3-3 proteins: Structure, function, and regulation. Annu. Rev. Pharmacol. 2000, 40, 617–647. [Google Scholar] [CrossRef] [PubMed]
  93. Malakshah, S.N.; Rezaei, M.; Heidari, M.; Salekdeh, G.H. Proteomics reveals new salt responsive proteins associated with rice plasma membrane. Biosci. Biotechnol. Biochem. 2014, 71, 2144–2154. [Google Scholar] [CrossRef] [PubMed]
  94. Chandna, R.; Augustine, R.; Kanchupati, P.; Kumar, R.; Kumar, P.; Arya, G.C.; Bisht, N.C. Class-specific evolution and transcriptional differentiation of 14-3-3 family members in mesohexaploid Brassica rapa. Front. Plant Sci. 2016, 7, 12. [Google Scholar] [CrossRef] [PubMed]
  95. Chen, F.; Li, Q.; Sun, L.; He, Z. The rice 14-3-3 gene family and its involvement in responses to biotic and abiotic stress. DNA Res. 2006, 13, 53–63. [Google Scholar] [CrossRef] [PubMed]
  96. Li, R.H.; Jiang, X.T.; Jin, D.H.; Dhaubhadel, S.; Bian, S.M.; Li, X.Y. Identification of 14-3-3 family in common bean and their response to abiotic stress. PLoS ONE 2015, 10, e0143280. [Google Scholar] [CrossRef] [PubMed]
  97. Sun, G.; Xie, F.; Zhang, B. Transcriptome-wide identification and stress properties of the 14-3-3 gene family in cotton (Gossypium hirsutum L.). Funct. Integr. Genom. 2011, 11, 627–636. [Google Scholar] [CrossRef] [PubMed]
  98. Sun, X.; Luo, X.; Sun, M.; Chen, C.; Ding, X.; Wang, X.; Yang, S.; Yu, Q.; Jia, B.; Ji, W. A Glycine Soja 14-3-3 Protein GsGF14o participates in stomatal and root hair development and drought tolerance in Arabidopsis thaliana. Plant Cell Physiol. 2014, 55, 99–118. [Google Scholar] [CrossRef] [PubMed]
  99. Skirycz, A.; de Bodt, S.; Obata, T.; de Clercq, I.; Claeys, H.; de Rycke, R.; Andriankaja, M.; van Aken, O.; van Breusegem, F.; Fernie, A.R. Developmental stage specificity and the role of mitochondrial metabolism in the response of Arabidopsis leaves to prolonged mild osmotic stress. Plant Physiol. 2010, 152, 226–244. [Google Scholar] [CrossRef] [PubMed]
  100. Verelst, W.; Skirycz, A.; Inze, D. Abscisic acid, ethylene and gibberellic acid act at different developmental stages to instruct the adaptation of young leaves to stress. Plant Signal. Behav. 2010, 5, 473–475. [Google Scholar] [CrossRef] [PubMed]
  101. Wasternack, C.; Parthier, B. Jasmonate-signalled plant gene expression. Trends Plant Sci. 1997, 2, 302–307. [Google Scholar] [CrossRef]
  102. Jin, L.G.; Liu, J.-Y. Molecular cloning, expression profile and promoter analysis of a novel ethylene responsive transcription factor gene GhERF4 from cotton (Gossypium hirstum). Plant Physiol. Biochem. 2008, 46, 46–53. [Google Scholar] [CrossRef] [PubMed]
  103. Jin, L.; Huang, B.; Li, H.; Liu, J. Expression profiles and transactivation analysis of a novel ethylene-responsive transcription factor gene GhERF5 from cotton. Prog. Nat. Sci. 2009, 19, 563–572. [Google Scholar] [CrossRef]
  104. Zhang, W.; Yang, G.; Mu, D.; Li, H.; Zang, D.; Xu, H.; Zou, X.; Wang, Y. An ethylene-responsive factor BpERF11 negatively modulates salt and osmotic tolerance in Betula platyphylla. Sci. Rep. 2016, 6, 23085. [Google Scholar] [CrossRef] [PubMed]
  105. Shinozaki, K.; Yamaguchi-Shinozaki, K. Gene networks involved in drought stress response and tolerance. J. Exp. Bot. 2007, 58, 221–227. [Google Scholar] [CrossRef] [PubMed]
  106. Trivedi, I.; Rai, K.M.; Singh, S.K.; Kumar, V.; Singh, M.; Ranjan, A.; Lodhi, N.; Sawant, S.V. Analysis of histones and histone variants in plants. In Chromatin Remodeling: Metthods and Protocols; Morse, R.H., Ed.; Springer: Berlin/Heidelberg, Germany, 2012; pp. 225–236. [Google Scholar]
  107. Trivedi, I.; Ranjan, A.; Sharma, Y.K.; Sawant, S. The histone H1 variant accumulates in response to water stress in the drought tolerant genotype of Gossypium herbaceum L. Protein J. 2012, 31, 477–486. [Google Scholar] [CrossRef] [PubMed]
  108. Lange, S.; Mitchell, D.L.; Vasquez, K.M. High mobility group protein B1 enhances DNA repair and chromatin modification after DNA damage. Proc. Natl. Acad. Sci. USA 2008, 105, 10320. [Google Scholar] [CrossRef] [PubMed]
  109. Zhao, J.; Paul, L.K.; Grafi, G. The maize HMGA protein is localized to the nucleolus and can be acetylated in vitro at its globular domain, and phosphorylation by CDK reduces its binding activity to AT-rich DNA. BBA Gene Regul. Mech. 2009, 1789, 751–757. [Google Scholar] [CrossRef] [PubMed]
  110. Singh, U.; Deb, D.; Singh, A.; Grover, A. Glycine-rich RNA binding protein of Oryza sativa inhibits growth of M15 E. coli cells. BMC Res. Notes 2011, 4, 18. [Google Scholar]
  111. Yang, D.H.; Kwak, K.J.; Kim, M.K.; Park, S.J.; Yang, K.Y.; Kang, H. Expression of Arabidopsis glycine-rich RNA-binding protein AtGRP2 or AtGRP7 improves grain yield of rice (Oryza sativa) under drought stress conditions. Plant Sci. 2014, 214, 106–112. [Google Scholar] [CrossRef] [PubMed]
  112. Kwak, K.J.; Kim, H.S.; Jang, H.Y.; Kang, H.; Ahn, S.J. Diverse roles of glycine-rich RNA-binding protein 7 in the response of camelina (Camelina sativa) to abiotic stress. Acta Physiol. Plant. 2016, 38, 129. [Google Scholar] [CrossRef]
  113. Zheng, J.; Wang, Y.Y.; He, Y.A.; Zhou, J.J.; Li, Y.P.; Liu, Q.Q.; Xie, X.Z. Overexpression of an S-like ribonuclease gene, OsRNS4, confers enhanced tolerance to high salinity and hyposensitivity to phytochrome-mediated light signals in rice. Plant Sci. 2014, 214, 99–105. [Google Scholar] [CrossRef] [PubMed]
  114. Timperio, A.M.; Egidi, M.G.; Zolla, L. Proteomics applied on plant abiotic stresses: Role of heat shock proteins (HSP). J. Proteom. 2008, 71, 391–411. [Google Scholar] [CrossRef] [PubMed]
  115. Swindell, W.R.; Huebner, M.; Weber, A.P. Transcriptional profiling of Arabidopsis heat shock proteins and transcription factors reveals extensive overlap between heat and non-heat stress response pathways. BMC Genom. 2007, 8, 125. [Google Scholar] [CrossRef] [PubMed]
  116. Sabehat, A.; Lurie, S.; Weiss, D. Expression of small heat-shock proteins at low temperatures. A possible role in protecting against chilling injuries. Plant Physiol. 1998, 117, 651–658. [Google Scholar] [CrossRef] [PubMed]
  117. Gao, C.Q.; Jiang, B.; Wang, Y.C.; Liu, G.F.; Yang, C.P. Overexpression of a heat shock protein (ThHSP18.3) from Tamarix hispida confers stress tolerance to yeast. Mol. Biol. Rep. 2012, 39, 4889–4897. [Google Scholar] [CrossRef] [PubMed]
  118. Masand, S.; Yadav, S.K. Overexpression of MuHSP70 gene from Macrotyloma uniflorum confers multiple abiotic stress tolerance in transgenic Arabidopsis thaliana. Mol. Biol. Rep. 2016, 43, 53–64. [Google Scholar] [CrossRef] [PubMed]
  119. Guo, Q.F.; Zhang, J.; Gao, Q.; Xing, S.C.; Li, F.; Wang, W. Drought tolerance through overexpression of monoubiquitin in transgenic tobacco. J. Plant Physiol. 2008, 165, 1745–1755. [Google Scholar] [CrossRef] [PubMed]
  120. Yao, X.; Xiong, W.; Ye, T.T.; Wu, Y. Overexpression of the aspartic protease ASPG1 gene confers drought avoidance in Arabidopsis. J. Exp. Bot. 2012, 63, 2579–2593. [Google Scholar] [CrossRef] [PubMed]
  121. Simova-Stoilova, L.; Vaseva, I.; Grigorova, B.; Demirevska, K.; Feller, U. Proteolytic activity and cysteine protease expression in wheat leaves under severe soil drought and recovery. Plant Physiol. Biochem. 2010, 48, 200–206. [Google Scholar] [CrossRef] [PubMed]
  122. Catala, R.; Ouyang, J.; Abreu, I.A.; Hu, Y.; Seo, H.; Zhang, X.; Chua, N.H. The Arabidopsis E3 SUMO ligase SIZ1 regulates plant growth and drought responses. Plant Cell 2007, 19, 2952–2966. [Google Scholar] [CrossRef] [PubMed]
  123. Miazek, A.; Zagdańska, B. Involvement of exopeptidases in dehydration tolerance of spring wheat seedlings. Biol. Plant. 2008, 52, 687–694. [Google Scholar] [CrossRef]
  124. Seo, Y.S.; Choi, J.Y.; Kim, S.J.; Kim, E.Y.; Shin, J.S.; Kim, W.T. Constitutive expression of CaRma1H1, a hot pepper ER-localized RING E3 ubiquitin ligase, increases tolerance to drought and salt stresses in transgenic tomato plants. Plant Cell Rep. 2012, 31, 1659–1665. [Google Scholar] [CrossRef] [PubMed]
  125. Kim, S.J.; Kim, W.T. Suppression of Arabidopsis RING E3 ubiquitin ligase AtATL78 increases tolerance to cold stress and decreases tolerance to drought stress. FEBS Lett. 2013, 587, 2584–2590. [Google Scholar] [CrossRef] [PubMed]
  126. Bae, H.; Kim, S.K.; Cho, S.K.; Kang, B.G.; Kim, W.T. Overexpression of OsRDCP1, a rice RING domain-containing E3 ubiquitin ligase, increased tolerance to drought stress in rice (Oryza sativa L.). Plant Sci. 2011, 180, 775–782. [Google Scholar] [CrossRef] [PubMed]
  127. Mckersie, B.D.; Bowley, S.R.; Harjanto, E.; Leprince, O. Water-deficit tolerance and field performance of transgenic alfalfa overexpressing superoxide dismutase. Plant Physiol. 1996, 111, 1177–1181. [Google Scholar] [CrossRef] [PubMed]
  128. Prashanth, S.; Sadhasivam, V.; Parida, A. Over expression of cytosolic copper/zinc superoxide dismutase from a mangrove plant Avicennia marina in indica rice var Pusa Basmati-1 confers abiotic stress tolerance. Transgenic Res. 2007, 17, 281–291. [Google Scholar] [CrossRef] [PubMed]
  129. Sofo, A.; Scopa, A.; Nuzzaci, M.; Vitti, A. Ascorbate peroxidase and catalase activities and their genetic regulation in plants subjected to drought and salinity stresses. Int. J. Mol. Sci. 2015, 16, 13561–13578. [Google Scholar] [CrossRef] [PubMed]
  130. Jung, S. Variation in antioxidant metabolism of young and mature leaves of Arabidopsis thaliana subjected to drought. Plant Sci. 2004, 166, 459–466. [Google Scholar] [CrossRef]
  131. Veljovic-Jovanovic, S.; Kukavica, B.; Stevanovic, B.; Navari-Izzo, F. Senescence- and drought-related changes in peroxidase and superoxide dismutase isoforms in leaves of Ramonda serbica. J. Exp. Bot. 2006, 57, 1759–1768. [Google Scholar] [CrossRef] [PubMed]
  132. Wang, W.B.; Kim, Y.H.; Lee, H.S.; Kim, K.Y.; Deng, X.P.; Kwak, S.S. Analysis of antioxidant enzyme activity during germination of alfalfa under salt and drought stresses. Plant Physiol. Biochem. 2009, 47, 570–577. [Google Scholar] [CrossRef] [PubMed]
  133. Gao, C.Q.; Wang, Y.C.; Liu, G.F.; Wang, C.; Jiang, J.; Yang, C.P. Cloning of ten peroxidase (POD) genes from Tamarix hispida and characterization of their responses to abiotic stress. Plant Mol. Biol. Rep. 2010, 28, 77–89. [Google Scholar] [CrossRef]
  134. Foyer, C.H.; Noctor, G. Ascorbate and glutathione: The heart of the redox hub. Plant Physiol. 2011, 155, 2–18. [Google Scholar] [CrossRef] [PubMed]
  135. Noctor, G.; Foyer, C.H. Ascorbate and glutathione: Keeping active oxygen under control. Annu. Rev. Plant Biol. 2003, 49, 249–279. [Google Scholar] [CrossRef] [PubMed]
  136. Miller, G.; Suzuki, N.; Ciftciyilmaz, S.; Mittler, R. Reactive oxygen species homeostasis and signalling during drought and salinity stresses. Plant Cell Environ. 2010, 33, 453–467. [Google Scholar] [CrossRef] [PubMed]
  137. Selote, D.S.; Khanna-Chopra, R. Antioxidant response of wheat roots to drought acclimation. Protoplasma 2010, 245, 153–163. [Google Scholar] [PubMed]
  138. Zhang, Z.Y.; Zhang, Q.; Wu, J.; Zheng, X.L.; Zheng, S.; Sun, X.; Qiu, Q.; Lu, T. Gene knockout study reveals that cytosolic ascorbate peroxidase 2 (OsAPX2) plays a critical role in growth and reproduction in rice under drought, salt and cold stresses. PLoS ONE 2013, 8, e57472. [Google Scholar] [CrossRef] [PubMed]
  139. Contour-Ansel, D.; Torres-Franklin, M.L.; Cruz de Carvalho, M.H.; D’Arcy-Lameta, A. Glutathione reductase in leaves of cowpea: Cloning of two cDNAs, expression and enzymatic activity under progressive drought stress, desiccation and abscisic acid treatment. Ann. Bot. 2006, 98, 1279–1287. [Google Scholar] [CrossRef] [PubMed]
  140. Torres-Franklin, M.L.; Contour-Ansel, D.; Zuily-Fodil, Y.; Pham-Thi, A.-T. Molecular cloning of glutathione reductase cDNAs and analysis of GR gene expression in cowpea and common bean leaves during recovery from moderate drought stress. J. Plant Physiol. 2008, 165, 514–521. [Google Scholar] [CrossRef] [PubMed]
  141. Kim, I.S.; Shin, S.Y.; Kim, Y.S.; Kim, H.Y.; Yoon, H.S. Expression of a glutathione reductase from Brassica rapa subsp pekinensis enhanced cellular redox homeostasis by modulating antioxidant proteins in Escherichia coli. Mol. Cells 2009, 28, 479–487. [Google Scholar] [CrossRef] [PubMed]
  142. Pang, C.-H.; Wang, B.-S. Role of ascorbate peroxidase and glutathione reductase in ascorbate–glutathione cycle and stress tolerance in plants. In Ascorbate-Glutathione Pathway and Stress Tolerance in Plants; Anjum, A.N., Chan, M.-T., Umar, S., Eds.; Springer: Dordrecht, The Netherlands, 2010; pp. 91–113. [Google Scholar]
  143. Kang, G.Z.; Li, G.Z.; Liu, G.Q.; Xu, W.; Peng, X.Q.; Wang, C.Y.; Zhu, Y.J.; Guo, T.C. Exogenous salicylic acid enhances wheat drought tolerance by influence on the expression of genes related to ascorbate-glutathione cycle. Biol. Plant. 2013, 57, 718–724. [Google Scholar] [CrossRef]
  144. Osipova, S.V.; Permyakov, A.V.; Permyakova, M.D.; Pshenichnikova, T.A.; Borner, A. Leaf dehydroascorbate reductase and catalase activity is associated with soil drought tolerance in bread wheat. Acta Physiol. Plant. 2011, 33, 2169–2177. [Google Scholar] [CrossRef]
  145. Eltayeb, A.E.; Yamamoto, S.; Habora, M.E.E.; Yin, L.N.; Tsujimoto, H.; Tanaka, K. Transgenic potato overexpressing Arabidopsis cytosolic AtDHAR1 showed higher tolerance to herbicide, drought and salt stresses. Breed. Sci. 2011, 61, 3–10. [Google Scholar] [CrossRef]
  146. Kim, Y.S.; Kim, I.S.; Bae, M.J.; Choe, Y.H.; Kim, Y.H.; Park, H.M.; Kang, H.G.; Yoon, H.S. Homologous expression of cytosolic dehydroascorbate reductase increases grain yield and biomass under paddy field conditions in transgenic rice (Oryza sativa L. japonica). Planta 2013, 237, 1613–1625. [Google Scholar] [CrossRef] [PubMed]
  147. Eltayeb, A.E.; Kawano, N.; Badawi, G.H.; Kaminaka, H.; Sanekata, T.; Morishima, I.; Shibahara, T.; Inanaga, S.; Tanaka, K. Enhanced tolerance to ozone and drought stresses in transgenic tobacco overexpressing dehydroascorbate reductase in cytosol. Physiol. Plant. 2006, 127, 57–65. [Google Scholar] [CrossRef]
  148. Le, D.T.; Tarrago, L.; Watanabe, Y.; Kaya, A.; Lee, B.C.; Tran, U.; Nishiyama, R.; Fomenko, D.E.; Gladyshev, V.N.; Tran, L.S.P. Diversity of plant methionine sulfoxide reductases B and evolution of a form specific for free methionine sulfoxide. PLoS ONE 2013, 8, e65637. [Google Scholar] [CrossRef] [PubMed]
  149. Rouhier, N.; Dos Santos, C.V.; Tarrago, L.; Rey, P. Plant methionine sulfoxide reductase A and B multigenic families. Photosynth. Res. 2006, 89, 247–262. [Google Scholar] [CrossRef] [PubMed]
  150. Foyer, C.H.; Shigeoka, S. Understanding oxidative stress and antioxidant functions to enhance photosynthesis. Plant Physiol. 2011, 155, 93–100. [Google Scholar] [CrossRef] [PubMed]
  151. Liu, D.; Liu, Y.; Rao, J.; Wang, G.; Li, H.; Ge, F.; Chen, C. Overexpression of the glutathione S-transferase gene from Pyrus pyrifolia fruit improves tolerance to abiotic stress in transgenic tobacco plants. Mol. Biol. 2013, 47, 515–523. [Google Scholar] [CrossRef]
  152. Ji, W.; Zhu, Y.; Li, Y.; Yang, L.; Zhao, X.; Cai, H.; Bai, X. Over-expression of a glutathione S-transferase gene, GsGST, from wild soybean (Glycine soja) enhances drought and salt tolerance in transgenic tobacco. Biotechnol. Lett. 2010, 32, 1173–1179. [Google Scholar] [CrossRef] [PubMed]
  153. Xu, J.; Xing, X.J.; Tian, Y.S.; Peng, R.H.; Xue, Y.; Zhao, W.; Yao, Q.H. Transgenic Arabidopsis plants expressing tomato glutathione S-transferase showed enhanced resistance to salt and drought stress. PLoS ONE 2015, 10, e0136960. [Google Scholar] [CrossRef] [PubMed]
  154. Islam, T.; Manna, M.; Reddy, M.K. Glutathione peroxidase of Pennisetum glaucum (PgGPx) is a functional Cd2+ dependent peroxiredoxin that enhances tolerance against salinity and drought stress. PLoS ONE 2015, 10, e0143344. [Google Scholar] [CrossRef] [PubMed]
  155. Renner, T.; Specht, C.D. Molecular and functional evolution of class I chitinases for plant carnivory in the caryophyllales. Mol. Biol. Evol. 2012, 29, 2971–2985. [Google Scholar] [CrossRef] [PubMed]
  156. Samac, D.A.; Smigocki, A.C. Expression of oryzacystatin I and II in alfalfa increases resistance to the root-lesion nematode. Phytopathology 2003, 93, 799–804. [Google Scholar] [CrossRef] [PubMed]
  157. Demirevska, K.; Simova-Stoilova, L.; Fedina, I.; Georgieva, K.; Kunert, K. Response of oryzacystatin I transformed tobacco plants to drought, heat and light stress. J. Agron. Crop Sci. 2010, 196, 90–99. [Google Scholar] [CrossRef]
  158. Jain, S.; Kumar, D.; Jain, M.; Chaudhary, P.; Deswal, R.; Sarin, N.B. Ectopic overexpression of a salt stress-induced pathogenesis-related class 10 protein (PR10) gene from peanut (Arachis hypogaea L.) affords broad spectrum abiotic stress tolerance in transgenic tobacco. Plant Cell Tissue Organ. 2012, 109, 19–31. [Google Scholar] [CrossRef]
  159. Chakrabortee, S.; Boschetti, C.; Walton, L.J.; Sarkar, S.; Rubinsztein, D.C.; Tunnacliffe, A. Hydrophilic protein associated with desiccation tolerance exhibits broad protein stabilization function. Proc. Natl. Acad. Sci. USA 2007, 104, 18073–18078. [Google Scholar] [CrossRef] [PubMed]
  160. Hu, L.; Wang, Z.; Do, H.; Huang, B. Differential accumulation of dehydrins in response to water stress for hybrid and common bermudagrass genotypes differing in drought tolerance. J. Plant Physiol. 2010, 167, 103–109. [Google Scholar] [CrossRef] [PubMed]
  161. Jangpromma, N.; Kitthaisong, S.; Lomthaisong, K.; Daduang, S.; Jaisil, P.; Thammasirirak, S. A proteomics analysis of drought stress-responsive proteins as biomarker for drought-tolerant sugarcane cultivars. Am. J. Biochem. Biotechnol. 2010, 6, 89–102. [Google Scholar] [CrossRef]
  162. Alsheikh, M.; Heyen, B.J.; Randall, S.K. Ion binding properties of the dehydrin ERD14 are dependent upon phosphorylation. J. Biol. Chem. 2003, 278, 40882–40889. [Google Scholar] [CrossRef] [PubMed]
  163. Park, S.C.; Kim, Y.H.; Jeong, J.C.; Kim, C.Y.; Lee, H.S.; Bang, J.W.; Kwak, S.-S. Sweetpotato late embryogenesis abundant 14 (IbLEA14) gene influences lignification and increases osmotic- and salt stress-tolerance of transgenic calli. Planta 2011, 233, 621–634. [Google Scholar] [CrossRef] [PubMed]
  164. Ashraf, M.; Foolad, M.R. Roles of glycine betaine and proline in improving plant abiotic stress resistance. Environ. Exp. Bot. 2007, 59, 206–216. [Google Scholar] [CrossRef]
  165. Tang, W.; Sun, J.; Liu, J.; Liu, F.; Yan, J.; Gou, X.; Lu, B.-R.; Liu, Y. RNAi-directed downregulation of betaine aldehyde dehydrogenase 1 (OsBADH1) results in decreased stress tolerance and increased oxidative markers without affecting glycine betaine biosynthesis in rice (Oryza sativa). Plant Mol. Biol. 2014, 86, 443–454. [Google Scholar] [CrossRef] [PubMed]
  166. Skirycz, A.; Inze, D. More from less: Plant growth under limited water. Curr. Opin. Biotechnol. 2010, 21, 197–203. [Google Scholar] [CrossRef] [PubMed]
  167. Tardieu, F.; Granier, C.; Muller, B. Water deficit and growth. Co-ordinating processes without an orchestrator? Curr. Opin. Plant Biol. 2011, 14, 283–289. [Google Scholar] [CrossRef] [PubMed]
  168. Sniegowska-Swierk, K.; Dubas, E.; Rapacz, M. Drought-induced changes in the actin cytoskeleton of barley (Hordeum vulgare L.) leaves. Acta Physiol. Plant. 2015, 37. [Google Scholar] [CrossRef]
  169. Bommer, U.; Thiele, B. The translationally controlled tumour protein (TCTP). Int. J. Biochem. Cell B 2004, 36, 379–385. [Google Scholar] [CrossRef]
  170. Kim, Y.M.; Han, Y.J.; Hwang, O.J.; Lee, S.-S.; Shin, A.Y.; Kim, S.Y.; Kim, J.I. Overexpression of Arabidopsis translationally controlled tumor protein gene AtTCTP enhances drought tolerance with rapid ABA-induced stomatal closure. Mol. Cells 2012, 33, 617–626. [Google Scholar] [CrossRef] [PubMed]
  171. Berkowitz, O.; Jost, R.; Pollmann, S.; Masle, J. Characterization of TCTP, the translationally controlled tumor protein, from Arabidopsis thaliana. Plant Cell 2008, 20, 3430–3447. [Google Scholar] [CrossRef] [PubMed]
  172. Baxter, H.L.; Stewart, C.N. Effects of altered lignin biosynthesis on phenylpropanoid metabolism and plant stress. Biofuels 2013, 4, 635–650. [Google Scholar] [CrossRef]
  173. Lee, B.R.; Kim, K.Y.; Jung, W.J.; Avice, J.C.; Ourry, A.; Kim, T.H. Peroxidases and lignification in relation to the intensity of water-deficit stress in white clover (Trifolium repens L.). J. Exp. Bot. 2007, 58, 1271–1279. [Google Scholar] [CrossRef] [PubMed]
  174. Amor, Y.; Haigler, C.H.; Johnson, S.; Wainscott, M.; Delmer, D.P. A membrane-associated form of sucrose synthase and its potential role in synthesis of cellulose and callose in plants. Proc. Natl. Acad. Sci. USA 1995, 92, 9353–9357. [Google Scholar] [CrossRef] [PubMed]
  175. Dong, J.L.; Jiang, Y.Y.; Chen, R.J.; Xu, Z.J.; Gao, X.L. Isolation of a novel xyloglucan endotransglucosylase (OsXET9) gene from rice and analysis of the response of this gene to abiotic stresses. Afr. J. Biotechnol. 2011, 10, 17424–17434. [Google Scholar]
  176. Nishiyama, R.; Le, D.T.; Watanabe, Y.; Matsui, A.; Tanaka, M.; Seki, M.; Yamaguchi-Shinozaki, K.; Shinozaki, K.; Tran, L.S.P. Transcriptome analyses of a salt-tolerant cytokinin-deficient mutant reveal differential regulation of salt stress response by cytokinin deficiency. PLoS ONE 2012, 7, e32124. [Google Scholar] [CrossRef] [PubMed]
  177. Tenhaken, R. Cell wall remodeling under abiotic stress. Front. Plant Sci. 2015, 5, 771. [Google Scholar] [CrossRef] [PubMed]
  178. Spagnoletta, A.; De Santis, A.; Tampieri, E.; Baraldi, E.; Bachi, A.; Genchi, G. Identification and kinetic characterization of HtDTC, the mitochondrial dicarboxylate-tricarboxylate carrier of Jerusalem artichoke tubers. J. Bioenerg. Biomembr. 2006, 38, 57–65. [Google Scholar] [CrossRef] [PubMed]
  179. Wang, J.; Zhang, L.D.; Zuo, K.J.; Oian, H.M.; Cao, Y.F.; Tang, K.X. Cloning and expressional studies of the voltage-dependent anion channel gene from Brassica rapa L. J. Integr. Plant Biol. 2006, 48, 197–203. [Google Scholar] [CrossRef]
  180. Wen, G.; Yang, Y.; Chai, L.; Liu, Z.; Wang, J. Overexpression of the voltage-dependent anion channel 2 (VDAC2) gene induces drought resistance in Arabidopsis thaliana. Plant Omics 2014, 7, 171–177. [Google Scholar]
  181. Lefebvre, B.; Timmers, T.; Mbengue, M.; Moreau, S.; Herve, C.; Toth, K.; Bittencourt-Silvestre, J.; Klaus, D.; Deslandes, L.; Godiard, L.; et al. A remorin protein interacts with symbiotic receptors and regulates bacterial infection. Proc. Natl. Acad. Sci. USA 2010, 107, 2343–2348. [Google Scholar] [CrossRef] [PubMed]
  182. Checker, V.G.; Khurana, P. Molecular and functional characterization of mulberry EST encoding remorin (MiREM) involved in abiotic stress. Plant Cell Rep. 2013, 32, 1729–1741. [Google Scholar] [CrossRef] [PubMed]
  183. Zhou, S.; Hu, W.; Deng, X.; Ma, Z.; Chen, L.; Huang, C.; Wang, C.; Wang, J.; He, Y.; Yang, G.; et al. Overexpression of the wheat aquaporin gene, TaAQP7, enhances drought tolerance in transgenic tobacco. PLoS ONE 2012, 7, e52439. [Google Scholar] [CrossRef] [PubMed]
  184. Tapia, G.; Morales-Quintana, L.; Parra, C.; Berbel, A.; Alcorta, M. Study of nsLTPs in Lotus japonicus genome reveal a specific epidermal cell member (LjLTP10) regulated by drought stress in aerial organs with a putative role in cutin formation. Plant Mol. Biol. 2013, 82, 485–501. [Google Scholar] [CrossRef] [PubMed]
  185. Guo, L.; Yang, H.; Zhang, X.; Yang, S. Lipid transfer protein 3 as a target of MYB96 mediates freezing and drought stress in Arabidopsis. J. Exp. Bot. 2013, 64, 1755–1767. [Google Scholar] [CrossRef] [PubMed]
  186. Guo, C.; Ge, X.; Ma, H. The rice OsDIL gene plays a role in drought tolerance at vegetative and reproductive stages. Plant Mol. Biol. 2013, 82, 239–253. [Google Scholar] [CrossRef] [PubMed]
  187. Kabala, K.; Janicka-Russak, M.; Reda, M.; Migocka, M. Transcriptional regulation of the V-ATPase subunit c and V-PPase isoforms in Cucumis sativus under heavy metal stress. Physiol. Plant. 2014, 150, 32–45. [Google Scholar] [CrossRef] [PubMed]
  188. Fukuda, A.; Tanaka, Y. Effects of ABA, auxin, and gibberellin on the expression of genes for vacuolar H+-inorganic pyrophosphatase, H+-ATPase subunit A, and Na+/H+ antiporter in barley. Plant Physiol. Biochem. 2006, 44, 351–358. [Google Scholar] [CrossRef] [PubMed]
  189. Dong, Q.L.; Wang, C.R.; Liu, D.D.; Hu, D.G.; Fang, M.J.; You, C.X.; Yao, Y.X.; Hao, Y.J. MdVHA-A encodes an apple subunit A of vacuolar H+-ATPase and enhances drought tolerance in transgenic tobacco seedlings. J. Plant Physiol. 2013, 170, 601–609. [Google Scholar] [CrossRef] [PubMed]
  190. Patro, L.; Mohapatra, P.K.; Biswal, U.C.; Biswal, B. Dehydration induced loss of photosynthesis in Arabidopsis leaves during senescence is accompanied by the reversible enhancement in the activity of cell wall beta-glucosidase. J. Photochem. Photobiol. B 2014, 137, 49–54. [Google Scholar] [CrossRef] [PubMed]
  191. Ghotbi-Ravandi, A.A.; Shahbazi, M.; Shariati, M.; Mulo, P. Effects of mild and severe drought stress on photosynthetic efficiency in tolerant and susceptible barley (Hordeum vulgare L.) genotypes. J. Agron. Crop Sci. 2014, 200, 403–415. [Google Scholar] [CrossRef]
  192. Xu, Y.; Liu, R.; Yan, L.; Liu, Z.; Jiang, S.; Shen, Y.; Wang, X.; Zhang, D. Light-harvesting chlorophyll a/b-binding proteins are required for stomatal response to abscisic acid in Arabidopsis. J. Exp. Bot. 2012, 63, 1095–1106. [Google Scholar] [CrossRef] [PubMed]
  193. Majeau, N.; Arnoldo, M.; Coleman, J.R. Modification of carbonic anhydrase activity by antisense and over-expression constructs in transgenic tobacco. Plant Mol. Biol. 1994, 25, 377–385. [Google Scholar] [CrossRef] [PubMed]
  194. Abebe, T.; Melmaiee, K.; Berg, V.; Wise, R.P. Drought response in the spikes of barley: Gene expression in the lemma, palea, awn, and seed. Funct. Integr. Genom. 2010, 10, 191–205. [Google Scholar] [CrossRef] [PubMed]
  195. Voss, I.; Sunil, B.; Scheibe, R.; Raghavendra, A.S. Emerging concept for the role of photorespiration as an important part of abiotic stress response. Plant Biol. 2013, 15, 713–722. [Google Scholar] [CrossRef] [PubMed]
  196. Degenkolbe, T.; Do, P.T.; Kopka, J.; Zuther, E.; Hincha, D.K.; Koehl, K.I. Identification of drought tolerance markers in a diverse population of rice cultivars by expression and metabolite profiling. PLoS ONE 2013, 8, e63637. [Google Scholar] [CrossRef] [PubMed]
  197. Kappachery, S.; Baniekal-Hiremath, G.; Yu, J.W.; Park, S.W. Effect of over-and under-expression of glyceraldehyde 3-phosphate dehydrogenase on tolerance of plants to water-deficit stress. Plant Cell Tissue Organ. 2015, 121, 97–107. [Google Scholar] [CrossRef]
  198. Xu, Z.Z.; Zhou, G.S. Nitrogen metabolism and photosynthesis in Leymus chinensis in response to long-term soil drought. J. Plant Growth Regul. 2006, 25, 252–266. [Google Scholar] [CrossRef]
  199. Fresneau, C.; Ghashghaie, J.; Cornic, G. Drought effect on nitrate reductase and sucrose-phosphate synthase activities in wheat (Triticum durum L.): Role of leaf internal CO2. J. Exp. Bot. 2007, 58, 2983–2992. [Google Scholar] [PubMed]
  200. Sánchez-Rodríguez, E.; Rubio-Wilhelmi, M.d.M.; Ríos, J.J.; Blasco, B.; Rosales, M.Á.; Melgarejo, R.; Romero, L.; Ruiz, J.M. Ammonia production and assimilation: Its importance as a tolerance mechanism during moderate water deficit in tomato plants. J. Plant Physiol. 2011, 168, 816–823. [Google Scholar] [CrossRef] [PubMed]
  201. Goel, P.; Singh, A.K. Abiotic stresses downregulate key genes involved in nitrogen uptake and assimilation in Brassica juncea L. PLoS ONE 2015, 10, e0143645. [Google Scholar] [CrossRef] [PubMed]
  202. Jedmowski, C.; Ashoub, A.; Beckhaus, T.; Berberich, T.; Karas, M.; Bruggemann, W. Comparative analysis of Sorghum bicolor proteome in response to drought stress and following recovery. Int. J. Proteom. 2014, 2014, 395905. [Google Scholar] [CrossRef] [PubMed]
  203. Tabor, C.W.; Tabor, H. Methionine adenosyltransferase (S-Adenosylmethionine synthetase) and S-adenosylmethionine decarboxylase. In Advances in Enzymology and Related Areas of Molecular Biology; Meister, A., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2006; pp. 251–282. [Google Scholar]
  204. Boerjan, W.; Bauw, G.; Van Montagu, M.; Inze, D. Distinct phenotypes generated by overexpression and suppression of S-adenosyl-l-methionine synthetase reveal developmental patterns of gene silencing in tobacco. Plant Cell 1994, 6, 1401–1414. [Google Scholar] [CrossRef] [PubMed]
  205. Narita, Y.; Taguchi, H.; Nakamura, T.; Ueda, A.; Shi, W.; Takabe, T. Characterization of the salt-inducible methionine synthase from barley leaves. Plant Sci. 2004, 167, 1009–1016. [Google Scholar] [CrossRef]
  206. Kim, S.H.; Kim, S.H.; Palaniyandi, S.A.; Yang, S.H.; Suh, J.-W. Expression of potato S-adenosyl-l-methionine synthase (SbSAMS) gene altered developmental characteristics and stress responses in transgenic Arabidopsis plants. Plant Physiol. Biochem. 2015, 87, 84–91. [Google Scholar] [CrossRef] [PubMed]
  207. Gasulla, F.; vom Dorp, K.; Dombrink, I.; Zaehringer, U.; Gisch, N.; Doermann, P.; Bartels, D. The role of lipid metabolism in the acquisition of desiccation tolerance in Craterostigma plantagineum: A comparative approach. Plant J. 2013, 75, 726–741. [Google Scholar] [CrossRef] [PubMed]
  208. Benhassaine-Kesri, G.; Aid, F.; Demandre, C.; Kader, J.C.; Mazliak, P. Drought stress affects chloroplast lipid metabolism in rape (Brassica napus) leaves. Physiol. Plant. 2002, 115, 221–227. [Google Scholar] [CrossRef] [PubMed]
  209. Zhong, D.; Du, H.; Wang, Z.; Huang, B. Genotypic variation in fatty acid composition and unsaturation levels in bermudagrass associated with leaf dehydration tolerance. J. Am. Soc. Hortic. Sci. 2011, 136, 35–40. [Google Scholar]
  210. Gigon, A.; Matos, A.R.; Laffray, D.; Zuily-Fodil, Y.; Pham-Thi, A.T. Effect of drought stress on lipid metabolism in the leaves of Arabidopsis thaliana (ecotype Columbia). Ann. Bot. 2004, 94, 345–351. [Google Scholar] [CrossRef] [PubMed]
  211. Castellarin, S.D.; Matthews, M.A.; di Gaspero, G.; Gambetta, G.A. Water deficits accelerate ripening and induce changes in gene expression regulating flavonoid biosynthesis in grape berries. Planta 2007, 227, 101–112. [Google Scholar] [CrossRef] [PubMed]
  212. Liu, M.L.; Li, X.R.; Liu, Y.B.; Cao, B. Regulation of flavanone 3-hydroxylase gene involved in the flavonoid biosynthesis pathway in response to UV-B radiation and drought stress in the desert plant, Reaumuria soongorica. Plant Physiol. Biochem. 2013, 73, 161–167. [Google Scholar] [CrossRef] [PubMed]
  213. Meng, C.; Zhang, S.; Deng, Y.S.; Wang, G.D.; Kong, F.Y. Overexpression of a tomato flavanone 3-hydroxylase-like protein gene improves chilling tolerance in tobacco. Plant Physiol. Biochem. 2015, 96, 388–400. [Google Scholar] [CrossRef] [PubMed]
  214. Rani, A.; Singh, K.; Ahuja, P.S.; Kumar, S. Molecular regulation of catechins biosynthesis in tea (Camellia sinensis (L.) O. Kuntze). Gene 2012, 495, 205–210. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Functional categories of drought-responsive protein species and unique proteins in leaves revealed by proteomic studies. (A) Protein species include all the protein isoforms generated from gene variable splicing and post-translational modifications in the original publications; (B) Unique proteins indicate the protein family whose members have similar protein sequence homology and functional domains.
Figure 1. Functional categories of drought-responsive protein species and unique proteins in leaves revealed by proteomic studies. (A) Protein species include all the protein isoforms generated from gene variable splicing and post-translational modifications in the original publications; (B) Unique proteins indicate the protein family whose members have similar protein sequence homology and functional domains.
Ijms 17 01706 g001
Figure 2. Schematic representation of drought responsive proteins involved in signaling, gene expression regulation, protein synthesis and degradation. The solid line indicates single-step reactions, and the dashed line indicates multi-step reactions. (A) Signaling; (B) DNA replication and transcription; (C) RNA processing; (D) Protein synthesis; (E) Protein processing; (F) Protein degradation. 14-3-3: 14-3-3 protein; Nep: aspartic proteinase; ClpP: ATP-dependent Clp protease; ABP: auxin-binding protein; BiP: endoplasmic reticulum-luminal binding protein; CaB: calcium ion binding protein; CaSR: calcium sensing receptor; CDPK: calcium-dependent protein kinase; CAM: calmodulin; CNX: calnexin; CRT: calreticulin; CP: cysteine proteinase; EF: elongation factor; ERF: ethylene-responsive transcription factor; GDI: GDP dissociation inhibitor; GR-RBP: glycine-rich RNA binding protein; HSP: heat shock protein; HMG: high mobility group protein; HOP: Hsp70-Hsp90 organizing protein; AP: leucine aminopeptidase; MatK: maturase K; MEP: metalloendopeptidase; MC: molecular chaperone; OP: oligopeptidase A-like; PPIase: peptidyl-prolyl cis-trans isomerase; PLD: phospholipase D; PhyC: phytochrome C; PCNA: proliferating cell nuclear antigen; PIP: proline iminopeptidase, putative; PDI: protein disulphide isomerase; PP2C: protein phosphatase 2C; RanBP: Ran-binding protein; RT: retrotransposon protein; RNase: ribonuclease; RNP: ribonucleoprotein; RP: ribosomal protein; RBP: RNA binding protein; Hlc: RNA helicase; RNAP: RNA polymerase; SCPL: serine carboxypeptidase-like protein; STK: serine/threonine kinase; PSP: serine/threonine-protein phosphatase; SnRK: sucrose non-fermenting 1-related protein kinase; TCP: T-complex protein; TRIP1: TGF-β receptor-interacting protein 1; TF: transcription factor; TR: transcription regulator; TIG: trigger factor-like protein; V-PPase: vacuolar H+-pyrophosphatase; V-ATPase: vacuolar H+-ATPase; Zmp: zinc metalloprotease.
Figure 2. Schematic representation of drought responsive proteins involved in signaling, gene expression regulation, protein synthesis and degradation. The solid line indicates single-step reactions, and the dashed line indicates multi-step reactions. (A) Signaling; (B) DNA replication and transcription; (C) RNA processing; (D) Protein synthesis; (E) Protein processing; (F) Protein degradation. 14-3-3: 14-3-3 protein; Nep: aspartic proteinase; ClpP: ATP-dependent Clp protease; ABP: auxin-binding protein; BiP: endoplasmic reticulum-luminal binding protein; CaB: calcium ion binding protein; CaSR: calcium sensing receptor; CDPK: calcium-dependent protein kinase; CAM: calmodulin; CNX: calnexin; CRT: calreticulin; CP: cysteine proteinase; EF: elongation factor; ERF: ethylene-responsive transcription factor; GDI: GDP dissociation inhibitor; GR-RBP: glycine-rich RNA binding protein; HSP: heat shock protein; HMG: high mobility group protein; HOP: Hsp70-Hsp90 organizing protein; AP: leucine aminopeptidase; MatK: maturase K; MEP: metalloendopeptidase; MC: molecular chaperone; OP: oligopeptidase A-like; PPIase: peptidyl-prolyl cis-trans isomerase; PLD: phospholipase D; PhyC: phytochrome C; PCNA: proliferating cell nuclear antigen; PIP: proline iminopeptidase, putative; PDI: protein disulphide isomerase; PP2C: protein phosphatase 2C; RanBP: Ran-binding protein; RT: retrotransposon protein; RNase: ribonuclease; RNP: ribonucleoprotein; RP: ribosomal protein; RBP: RNA binding protein; Hlc: RNA helicase; RNAP: RNA polymerase; SCPL: serine carboxypeptidase-like protein; STK: serine/threonine kinase; PSP: serine/threonine-protein phosphatase; SnRK: sucrose non-fermenting 1-related protein kinase; TCP: T-complex protein; TRIP1: TGF-β receptor-interacting protein 1; TF: transcription factor; TR: transcription regulator; TIG: trigger factor-like protein; V-PPase: vacuolar H+-pyrophosphatase; V-ATPase: vacuolar H+-ATPase; Zmp: zinc metalloprotease.
Ijms 17 01706 g002
Figure 3. Schematic representation of drought responsive proteins involved in reactive oxygen species (ROS) scavenging, cell structure and cycle. The solid line indicates single-step reactions, and the dashed line indicates multi-step reactions. (A) ROS scavenging; (B) Cytoskeleton; (C) Cell cycle; (D) Cell wall modulation. ADF: actin depolymerizing factor; ALDH: aldehyde dehydrogenase; AsA: ascorbate; APX: ascorbate peroxidase; Exol: glucan exohydrolase; COMT: caffeic acid 3-O-methyltransferase; CCOMT: caffeoyl-CoA O-methyl-transferase; CAT: catalase; CDC: cell division cycle protein; DHAR: dehydroascorbate reductase; DHA: dehydroascorbate; GME: GDP-d-mannose-3′,5′-epimerase; GCL: glutamate-cysteine ligase; Grx: glutaredoxin; GPX: glutathione peroxidase; GR: glutathione reductase; GST: glutathione S-transferase; GSSG: glutathione disulfide; GSH: glutathione; GRP: glycine-rich protein; GLO: glyoxalase; LG: (R)-S-lactoylglutathione; MSR: methionine sulfoxide reductase; MG: methylglyoxal; MDHAR: monodehydroascorbate reductase; MDHA: monodehydroascorbate; ftsZ1: plastid division protein ftsZ1 precursor; PAE: pectin acetylesterase; PAL: phenylalanine ammonia-lyase; PG: pectin depolymerase; PGIP: polygalacturonase inhibitor; PHGPX: phospholipid hydroperoxide glutathione peroxidase; POD: peroxidase; Prx: peroxiredoxin; RGP: reversibly glycosylated polypeptide; SOD: superoxide dismutase; Trx: thioredoxin; TCTP: translationally-controlled tumor protein homolog; XI: xylanase inhibitor; XTH: xyloglucan endotransglycosylase.
Figure 3. Schematic representation of drought responsive proteins involved in reactive oxygen species (ROS) scavenging, cell structure and cycle. The solid line indicates single-step reactions, and the dashed line indicates multi-step reactions. (A) ROS scavenging; (B) Cytoskeleton; (C) Cell cycle; (D) Cell wall modulation. ADF: actin depolymerizing factor; ALDH: aldehyde dehydrogenase; AsA: ascorbate; APX: ascorbate peroxidase; Exol: glucan exohydrolase; COMT: caffeic acid 3-O-methyltransferase; CCOMT: caffeoyl-CoA O-methyl-transferase; CAT: catalase; CDC: cell division cycle protein; DHAR: dehydroascorbate reductase; DHA: dehydroascorbate; GME: GDP-d-mannose-3′,5′-epimerase; GCL: glutamate-cysteine ligase; Grx: glutaredoxin; GPX: glutathione peroxidase; GR: glutathione reductase; GST: glutathione S-transferase; GSSG: glutathione disulfide; GSH: glutathione; GRP: glycine-rich protein; GLO: glyoxalase; LG: (R)-S-lactoylglutathione; MSR: methionine sulfoxide reductase; MG: methylglyoxal; MDHAR: monodehydroascorbate reductase; MDHA: monodehydroascorbate; ftsZ1: plastid division protein ftsZ1 precursor; PAE: pectin acetylesterase; PAL: phenylalanine ammonia-lyase; PG: pectin depolymerase; PGIP: polygalacturonase inhibitor; PHGPX: phospholipid hydroperoxide glutathione peroxidase; POD: peroxidase; Prx: peroxiredoxin; RGP: reversibly glycosylated polypeptide; SOD: superoxide dismutase; Trx: thioredoxin; TCTP: translationally-controlled tumor protein homolog; XI: xylanase inhibitor; XTH: xyloglucan endotransglycosylase.
Ijms 17 01706 g003
Figure 4. Schematic model of drought responsive proteins involved in photosynthesis, carbon metabolism, fatty acid metabolism, and nitrogen and amino acid metabolism. The solid line indicates single-step reactions, and the dashed line indicates multi-step reactions. (A) Thylakoid reaction; (B) Calvin cycle; (C) Photorespiration; (D) Starch and sucrose metabolism; (E) Glycolysis; (F) Tricarboxylic acid (TCA) cycle; (G) Pentose phosphate pathway; (H) Nitrogen and amino acid metabolism; (I) Fatty acid metabolism. 1,3BPG: 1,3-bisphosphoglycerate; OMC: 2-oxoglutarate/malate carrier protein; MetE: 5-methyltetrahydropteroyltriglutamate-homocysteine methyltransferase; 6PG: 6-phosphogluconate; G6PDH: 6-phosphogluconate dehydrogenase; 6PGL: 6-phosphoglucono-δ-lactone; AHAS: acetohydroxyacid synthase; AH: aconitate hydratase; ACAD: acyl-CoA dehydrogenase; SAHase: adenosylhomocysteinase; ALAT: alanine aminotransferase; AGXT: alanine-glyoxylate aminotransferase; ADH: alcohol dehydrogenase; AldP: aldolase; AMT: aminomethyl transferase; AMY: amylase; AST: aspartate aminotransferase; ATPase: ATP synthase; AAC: ATP/ADP translocase; CA: carbonic anhydrase; FNR: chloroplast ferredoxin-NADP+ oxidoreductase; MSP: chloroplast manganese stabilizing protein; CS: cysteine synthase; Cyt b6f: cytochrome b6f complex; CCO: cytochrome c oxidase; Dic: dicarboxylate; DTC: dicarboxylate/tricarboxylate carrier; DLAT: dihydrolipoamide acetyltransferase; DHAP: dihydroxyacetone phosphate; ETCDB: electron carrier/iron ion binding protein; ENO: enolase; E4P: erythrose-4-phosphate; Fd: ferredoxin; F6P: fructose 6-phosphate; FBP: fructose-1,6-bisphosphate; FBPA: fructose-bisphosphate aldolase; G6P: glucose 6-phosphate; G1P: glucose-1-phosphate; GOGAT: glutamate synthase; GGAT: glutamate-glyoxylate aminotransferase; GS: glutamine synthetase; GAP: glyceraldehyde 3-phosphate; GAPDH: glyceraldehyde-3-phosphate dehydrogenase; GA: glyceric acid; GCS: glycine cleavage system; GDC: glycine decarboxylase; GO: glycolate oxidase; HCY: homocysteine; HPR: hydroxypyruvate reductase; IDH: isocitrate dehydrogenase; ICL: isocitrate lyase; LHCB: light-harvesting chlorophyll a/b-binding protein; LDH: lipoamide dehydrogenase; LDC: lysine decarboxylase; MDH: malate dehydrogenase; MS: methionine synthase; NDUFV: NADH dehydrogenase (ubiquinone) flavoprotein; NIR: nitrite reductase; OEC: oxygen-evolving complex protein; PEPCK: phosphoenol pyruvate carboxykinase; PEP: phosphoenolpyruvate; PGluM: phosphoglucomutase; PGA: phosphoglycerate; PGK: phosphoglycerate kinase; PGM: phosphoglycerate mutase; PRK: phosphoribulokinase; PSI: photosystem I reaction center protein; PSII: photosystem II reaction center protein; PC: plastocyanin; PDH: pyruvate dehydrogenase; DLDH: pyruvate dehydrogenase complex; PK: pyruvate kinase; R5P: ribose-5-phosphate; RPI: ribose-5-phosphate isomerase; Ru5P: ribulose 5-phosphate; Rubisco: ribulose-1,5-bisphosphate carboxylase/oxygenase; RCA: ribulose-1,5-bisphosphate carboxylase/oxygenase activase; RuBP: ribulose-1,5-disphosphate; RPE: ribulose-phosphate 3-epimerase; SAH: S-adenosyl-l-homocysteine; SAM: S-adenosylmethionine; SAMS: S-adenosylmethionine synthase; S7P: sedoheptulose 7-phosphate; SBPase: sedoheptulose-1,7-bisphosphatase; SBP: sedoheptulose-1,7-bisphosphate; SGAT: serine glyoxylate aminotransferase; SHMT: serine transhydroxymethyltransferase; SBE: starch branching enzyme; SS: starch synthase; SDH: succinate dehydrogenase; SCL: succinyl-CoA ligase; SUS: sucrose synthase; TLP: thylakoid lumen protein; TALDO: transaldolase; TKT: transketolase; Tric: tricarboxylate; TPI: triosephosphate isomerase; UDPG: uridine diphosphate glucose; UGPase: UTP-glucose pyrophosphorylase; VDAC: voltage dependent anion channel; X5P: xylulose 5-phosphate; α-KG: α-ketoglutarate.
Figure 4. Schematic model of drought responsive proteins involved in photosynthesis, carbon metabolism, fatty acid metabolism, and nitrogen and amino acid metabolism. The solid line indicates single-step reactions, and the dashed line indicates multi-step reactions. (A) Thylakoid reaction; (B) Calvin cycle; (C) Photorespiration; (D) Starch and sucrose metabolism; (E) Glycolysis; (F) Tricarboxylic acid (TCA) cycle; (G) Pentose phosphate pathway; (H) Nitrogen and amino acid metabolism; (I) Fatty acid metabolism. 1,3BPG: 1,3-bisphosphoglycerate; OMC: 2-oxoglutarate/malate carrier protein; MetE: 5-methyltetrahydropteroyltriglutamate-homocysteine methyltransferase; 6PG: 6-phosphogluconate; G6PDH: 6-phosphogluconate dehydrogenase; 6PGL: 6-phosphoglucono-δ-lactone; AHAS: acetohydroxyacid synthase; AH: aconitate hydratase; ACAD: acyl-CoA dehydrogenase; SAHase: adenosylhomocysteinase; ALAT: alanine aminotransferase; AGXT: alanine-glyoxylate aminotransferase; ADH: alcohol dehydrogenase; AldP: aldolase; AMT: aminomethyl transferase; AMY: amylase; AST: aspartate aminotransferase; ATPase: ATP synthase; AAC: ATP/ADP translocase; CA: carbonic anhydrase; FNR: chloroplast ferredoxin-NADP+ oxidoreductase; MSP: chloroplast manganese stabilizing protein; CS: cysteine synthase; Cyt b6f: cytochrome b6f complex; CCO: cytochrome c oxidase; Dic: dicarboxylate; DTC: dicarboxylate/tricarboxylate carrier; DLAT: dihydrolipoamide acetyltransferase; DHAP: dihydroxyacetone phosphate; ETCDB: electron carrier/iron ion binding protein; ENO: enolase; E4P: erythrose-4-phosphate; Fd: ferredoxin; F6P: fructose 6-phosphate; FBP: fructose-1,6-bisphosphate; FBPA: fructose-bisphosphate aldolase; G6P: glucose 6-phosphate; G1P: glucose-1-phosphate; GOGAT: glutamate synthase; GGAT: glutamate-glyoxylate aminotransferase; GS: glutamine synthetase; GAP: glyceraldehyde 3-phosphate; GAPDH: glyceraldehyde-3-phosphate dehydrogenase; GA: glyceric acid; GCS: glycine cleavage system; GDC: glycine decarboxylase; GO: glycolate oxidase; HCY: homocysteine; HPR: hydroxypyruvate reductase; IDH: isocitrate dehydrogenase; ICL: isocitrate lyase; LHCB: light-harvesting chlorophyll a/b-binding protein; LDH: lipoamide dehydrogenase; LDC: lysine decarboxylase; MDH: malate dehydrogenase; MS: methionine synthase; NDUFV: NADH dehydrogenase (ubiquinone) flavoprotein; NIR: nitrite reductase; OEC: oxygen-evolving complex protein; PEPCK: phosphoenol pyruvate carboxykinase; PEP: phosphoenolpyruvate; PGluM: phosphoglucomutase; PGA: phosphoglycerate; PGK: phosphoglycerate kinase; PGM: phosphoglycerate mutase; PRK: phosphoribulokinase; PSI: photosystem I reaction center protein; PSII: photosystem II reaction center protein; PC: plastocyanin; PDH: pyruvate dehydrogenase; DLDH: pyruvate dehydrogenase complex; PK: pyruvate kinase; R5P: ribose-5-phosphate; RPI: ribose-5-phosphate isomerase; Ru5P: ribulose 5-phosphate; Rubisco: ribulose-1,5-bisphosphate carboxylase/oxygenase; RCA: ribulose-1,5-bisphosphate carboxylase/oxygenase activase; RuBP: ribulose-1,5-disphosphate; RPE: ribulose-phosphate 3-epimerase; SAH: S-adenosyl-l-homocysteine; SAM: S-adenosylmethionine; SAMS: S-adenosylmethionine synthase; S7P: sedoheptulose 7-phosphate; SBPase: sedoheptulose-1,7-bisphosphatase; SBP: sedoheptulose-1,7-bisphosphate; SGAT: serine glyoxylate aminotransferase; SHMT: serine transhydroxymethyltransferase; SBE: starch branching enzyme; SS: starch synthase; SDH: succinate dehydrogenase; SCL: succinyl-CoA ligase; SUS: sucrose synthase; TLP: thylakoid lumen protein; TALDO: transaldolase; TKT: transketolase; Tric: tricarboxylate; TPI: triosephosphate isomerase; UDPG: uridine diphosphate glucose; UGPase: UTP-glucose pyrophosphorylase; VDAC: voltage dependent anion channel; X5P: xylulose 5-phosphate; α-KG: α-ketoglutarate.
Ijms 17 01706 g004
Figure 5. A summary of signaling and metabolic pathways in leaves in response to drought stress. Drought stress activates several signaling cascades, regulates gene expression, and promotes protein biosynthesis and turnover in drought-stressed leaves. Besides, drought stress inhibits photosynthesis, photorespiration, nitrogen and amino acid metabolism, as well as the cytoskeleton and cell division. Importantly, plants enhance ROS scavenging, osmotic regulation and cell wall remodulation. In addition, plants also modulate membrane trafficking, fatty acid metabolism, carbohydrate and energy metabolism, as well as secondary metabolism to cope with drought stress.
Figure 5. A summary of signaling and metabolic pathways in leaves in response to drought stress. Drought stress activates several signaling cascades, regulates gene expression, and promotes protein biosynthesis and turnover in drought-stressed leaves. Besides, drought stress inhibits photosynthesis, photorespiration, nitrogen and amino acid metabolism, as well as the cytoskeleton and cell division. Importantly, plants enhance ROS scavenging, osmotic regulation and cell wall remodulation. In addition, plants also modulate membrane trafficking, fatty acid metabolism, carbohydrate and energy metabolism, as well as secondary metabolism to cope with drought stress.
Ijms 17 01706 g005
Table 1. Summary of current publications on drought-responsive proteomics in leaves.
Table 1. Summary of current publications on drought-responsive proteomics in leaves.
Plant TypesNo.Plant SpeciesDrought Treatment ConditionProtein SpeciesUnique ProteinsReference
Herbs1Agrostis stolonifera35% soil water8157[18]
2Cynodon dactylonwithholding irrigation, 15 days5442[19]
3Elymus elongatum75%, 50%, 25% field capacity, rewater 3, 14 days119[20]
4Boea hygrometrica50% relative humidity (0.5, 8, 48 h); rehydrated 8, 48 h88[21]
5Poa pratensisTE foliar spray, 14 days; withholding irrigation, 0, 10 and 15 days5832[22]
withholding irrigation, 10 and 15 days8836[23]
6Sporobolus stapfianus30% relative water content10855[24]
7Hordeum spontaneumwithhold water to −2.5 MPa soil water potential, 3 weeks3222[25]
Hordeum vulgareP. indica-inoculation 28 days, 25% field capacity4528[26]
10% field capacity, 5 days3723[27]
stopping water, 7 days2423[28]
8Oryza sativa30% relative humidity, 48, 72, 96, 120, 144, 168 and 192 h9464[29]
withholding water, 38, 43 days; rewater 5, 10 days1713[30]
50% reposition of the water lost daily, 20 days1515[31]
stop irrigation, 35 days5331[32]
no irrigation, 20 days2018[33]
withholding water, 6 days1010[34] *
9Triticum aestivumwithout water 5, 14, 24 days; rewater 25 days159113[35]
1/3 of field capacity, 14 h1313[36]
PEG 6000, 48 h2319[37]
20%–25% soil relative water content, 3 days2916[38]
withholding water, 9 days3021[39]
20% PEG for 48 h31 29 [40] *
10Zea mays12.5% volumetric soil water content, 6 days220142[41]
withholding water, 9 days2927[42]
withholding water for 1, 4–5 days138111[43] *
PEG solution (−0.7 MPa) for 8 h14941[44] *
11Saccharum officinarum10% PEG 6000,14 h44[45]
12Musa paradisiaca0.21 M sorbitol, 48 days2420[46]
13Brassica napusno watering, 3, 7, 10 and 14 days417253[47]
14Citrullus lanatusno irrigation, 3 days2913[48]
15Glycine max10% PEG 6000 or exposed to drought conditions, 4 days5139[17]
16Phaseolus vulgarisno irrigation, 17 days8140[49]
17Medicago sativawater withholding, 7 days2920[50]
18Gossypium herbaceum35% relative water content1817[51]
Shrubs19Carissa spinarum25% relative humidity, 48, 120 h, rewater 24 h2312[52]
20Cistus albidusno irrigation, 107 days; rewater, 26 days5635[53]
21Hippophae rhamnoides25% field capacity, 20 days1311[54]
Trees22Eucalyptus sp.dry season, 6 months4619[55]
23Populus spp.25% field capacity, 18 days then rewater 28 days in glasshouse; no irrigation, 86 days in open field3329[56]
withholding water 4 and 7 days135[57]
withholding water 8, 12 days; rewater 7 days5225[58]
120 ppb ozone, 13 h; 35% soil water, 7, 14, 21 and 28 days2519[59]
10% soil water content, 45 days4033[60]
24Quercus robur15% soil water content 3, 8, 56 and 65 days4128[61]
Quercus ilexno irrigation 14 days; no irrigation 7 days, rewater 7 days1411[62]
not watering, 28 days1811[63]
25Malus domestica45%–55% field capacity8133[64]
The references labelled with * are phosphoproteomic studies. The information of these phosphoproteins is listed in Supplementary Table S2. TE: trinexapac-ethyl; PEG: polyethylene glycol.

Share and Cite

MDPI and ACS Style

Wang, X.; Cai, X.; Xu, C.; Wang, Q.; Dai, S. Drought-Responsive Mechanisms in Plant Leaves Revealed by Proteomics. Int. J. Mol. Sci. 2016, 17, 1706. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17101706

AMA Style

Wang X, Cai X, Xu C, Wang Q, Dai S. Drought-Responsive Mechanisms in Plant Leaves Revealed by Proteomics. International Journal of Molecular Sciences. 2016; 17(10):1706. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17101706

Chicago/Turabian Style

Wang, Xiaoli, Xiaofeng Cai, Chenxi Xu, Quanhua Wang, and Shaojun Dai. 2016. "Drought-Responsive Mechanisms in Plant Leaves Revealed by Proteomics" International Journal of Molecular Sciences 17, no. 10: 1706. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms17101706

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