Next Article in Journal
Optical Coherence Tomography Angiography of the Intestine: How to Prevent Motion Artifacts in Open and Laparoscopic Surgery?
Next Article in Special Issue
Current Epidemic Situation and Control Status of Citrus Huanglongbing in Guangdong China: The Space–Time Pattern Analysis of Specific Orchards
Previous Article in Journal
Plasmodium vivax MSP1-42 kD Variant Proteins Detected Naturally Induced IgG Antibodies in Patients Regardless of the Infecting Parasite Phenotype in Mesoamerica
Previous Article in Special Issue
Characterization of AtBAG2 as a Novel Molecular Chaperone
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Plant Metabolomics: An Overview of the Role of Primary and Secondary Metabolites against Different Environmental Stress Factors

1
Key Laboratory of Plant Ecology, Northeast Forestry Universit y, Harbin 150040, China
2
Horticulture Department, Faculty of Agriculture, Al-Azhar University, Nasr City, Cairo 11754, Egypt
3
Key Laboratory of Plant Nutrition and Agri-Environment in Northwest China, Ministry of Agriculture, College of Natural Resources and Environment, Northwest A&F University, Xianyang 712100, China
4
Laboratory of Phytochemistry, Department of Botany, University of São Paulo, São Paulo 05508-010, Brazil
5
Department of Horticulture and Life Science, Yeungnam University, Gyeongsan 38541, Republic of Korea
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Life 2023, 13(3), 706; https://doi.org/10.3390/life13030706
Submission received: 8 December 2022 / Revised: 2 January 2023 / Accepted: 28 February 2023 / Published: 6 March 2023
(This article belongs to the Special Issue Plant Biotic and Abiotic Stresses)

Abstract

:
Several environmental stresses, including biotic and abiotic factors, adversely affect the growth and development of crops, thereby lowering their yield. However, abiotic factors, e.g., drought, salinity, cold, heat, ultraviolet radiations (UVr), reactive oxygen species (ROS), trace metals (TM), and soil pH, are extremely destructive and decrease crop yield worldwide. It is expected that more than 50% of crop production losses are due to abiotic stresses. Moreover, these factors are responsible for physiological and biochemical changes in plants. The response of different plant species to such stresses is a complex phenomenon with individual features for several species. In addition, it has been shown that abiotic factors stimulate multi-gene responses by making modifications in the accumulation of the primary and secondary metabolites. Metabolomics is a promising way to interpret biotic and abiotic stress tolerance in plants. The study of metabolic profiling revealed different types of metabolites, e.g., amino acids, carbohydrates, phenols, polyamines, terpenes, etc, which are accumulated in plants. Among all, primary metabolites, such as amino acids, carbohydrates, lipids polyamines, and glycine betaine, are considered the major contributing factors that work as osmolytes and osmoprotectants for plants from various environmental stress factors. In contrast, plant-derived secondary metabolites, e.g., phenolics, terpenoids, and nitrogen-containing compounds (alkaloids), have no direct role in the growth and development of plants. Nevertheless, such metabolites could play a significant role as a defense by protecting plants from biotic factors such as herbivores, insects, and pathogens. In addition, they can enhance the resistance against abiotic factors. Therefore, metabolomics practices are becoming essential and influential in plants by identifying different phytochemicals that are part of the acclimation responses to various stimuli. Hence, an accurate metabolome analysis is important to understand the basics of stress physiology and biochemistry. This review provides insight into the current information related to the impact of biotic and abiotic factors on variations of various sets of metabolite levels and explores how primary and secondary metabolites help plants in response to these stresses.

1. Introduction

Comprehensively, biotic and abiotic stresses negatively affect crop production and cause a marked decrease in annual crop yield, i.e., qualitative and quantitative [1,2]. Recently, biologists, especially agriculturists, need to find an alternative way to deal with biotic and abiotic stresses such as herbivores, insects, and pathogens, as well as salinity, trace metals (TM) contamination, drought, and extreme temperatures [3,4] respectively. All these stresses affect the physiological and morphological aspects, such as the hindering of the functional groups of important molecules, e.g., enzymes, polynucleotides, transport systems for substantial ions and nutrients, as well as the growth and metabolic activities of plants [5,6]. However, to cope with these stresses, plants adopt several mechanisms, including metabolomics, transcriptomics, proteomics, and genomics, individually or in combination. The plant metabolome consists of the following two kinds of metabolites: primary and secondary metabolites. Primary metabolites are essential for the proper growth and development of plants and microorganisms. On the contrary, secondary metabolites are formed near the stationary phase of growth and have no direct role in growth, reproduction, and development. The metabolic profiling of primary and secondary metabolites provides extensive knowledge of biochemical processes that occurs in plant metabolism [7].
Modern research endorsed the purpose of several important genes, metabolites, proteins, and molecular systems that induced plant reactions to drought, salt stress, cold, TM, heat, and certain other biotic and abiotic factors [8,9]. Metabolomics analyses have become an influential tool to monitor plants’ responses to different environmentally stressed conditions [10]. Therefore, the findings of such studies give an understanding of the working of plants in definite circumstances, which are considered an important part of enlightening the molecular processes in responses to various stress conditions [11]. An appropriate data analysis, detection, identification, and evaluation of these metabolites are possible with the help of advanced metabolic tools such as gas chromatography-mass spectrometry (GC-MS), liquid chromatography-mass spectrometry (LC-MS) nuclear magnetic resonance (NMR) [12].
Furthermore, it is estimated that biotic and abiotic stresses are responsible for more than 50% of crop losses in the world [13,14]. The findings of Bayer in 2008 demonstrated that crop losses caused by abiotic stressors were significantly higher than by biotic factors [15]. However, the exact loss of crop yield depends on the plant’s developmental stage and the intensity and duration by which various stresses occur [16]. Among other stresses, salinity affects more than 800 million hectares of land—nearly 50% of the total irrigated area, which provides about 33% of the world’s food [17,18]. In the same way, drought also causes a loss of more than 50% of the average yield of crops [19]. Subsequently, other studies indicated that abiotic factors, such as temperature (low or high), salinity, and drought, significantly decreased plant production if existing alone or in combination [20]. Interestingly, another concern is the aggregation of reactive oxygen species (ROS), which is produced by excessively stressed accumulators of cadmium (Cd), chromium (Cr), lead (Pb), zinc (Zn), and copper (Cu) that can cause oxidation and dysfunction of biological molecules, hence disturbing certain physical and biological processes in plants [3,21]. Optimizing metabolic flux by the organellar electron transport chain (ETC) is essential in reducing oxidative stress [3]. Consequently, keeping the redox state of a cell is another essential issue that provides the decreasing power necessary for the foraging of ROS [22].
Therefore, there is a need for novel, easy, inexpensive, ecologically friendly, and robust crop types that can be conceived by cross-breeding or genetic engineering [23]. For example, recently, different wheat, rice, barley, maize, and other economically crucial varieties of crop plants have been considered very necessary than model plants [24,25]. However, the development of some modern ‘omics tools, such as genomics, proteomics, transcriptomics, and metabolomics, has rationalized the research of crop plants and abetted the complete study of acquaintances concerning biological components and plant breeding [26]. In this concern, metabolomics gives the possibility to accelerate the selection of superior breeding stock and the screening of elite crop types [27]. Primary and secondary metabolites, with their functional diversity, play an important role in fine-tuning the environmental stress tolerance and productivity in crops. Understanding plant behavior under multiple environmental stressors is one of the ways to deal with agricultural sustainability [20]. In this piece of work, more than 200 published works were considered to provide an overview of the role of primary and secondary metabolites against several abiotic and biotic stressors.

2. Instrumentation Applied in Metabolomics Studies

The identification of different classes of metabolites in plants is largely based on using hyphenated mass spectrometric methods to chromatographic equipment and electrophoretic approaches [28]. Choosing an appropriate ionization technique and analyzer type for metabolite analysis is important in a mass spectrometer [29]. Through the study of mass spectrometry (MS), ionized molecules are calculated. Similarly, mass-to-charge ratio values (m/z, m-mass, or z-charge) of the produced ions are assessed with the precision of one mass unit and to the fourth decimal point, small or high-resolution mass spectra, after elimination in the MS analyzer. The use of a high-resolution mass analyzer permits the accomplishment of the elemental composition of the identified ions existing in mass spectra. At first, it is probable to estimate the elemental composition and molecular mass of the molecules from enumerated m/z values for protonated [M + H]+ and deprotonated molecules [M − H]. The clear documentation of compounds is highly dependent on the applied MS system. MS machines designed with electrospray ionization (ESI) and matrix-assisted laser desorption or ionization (MALDI) source could be utilized. The ionization of MALDI can be joined to one or two unified times of flight analyzer (TOF and TOF/TOF). The source of ESI works well through quadrupole (Q), ion trap (IT), time of flight (TOF) analyzer, and a mixture of them. The maximum resolution in the mass analyzer could be attained by ion cyclotron resonance through Fourier Transformation Instruments (FT ICR MS) when the ESI is employed as an ionization system.
Moreover, designing the experiment according to the Metabolomics Standard Initiative (MSI) is also crucial, which endorses defined measures for the right biological materials preparation, procedures of metabolite extraction, and analytical protocols [30]. Following the regulations that have been stipulated, a sufficient number of sample replications and the conditions under which plant development should occur to be investigated and defined [31]. Similarly, the control of MS parameters in mass spectra registration is necessary. Such data deliver environments for suitable documentation and quantification of metabolites and consistent statistical quantification [32].
After employing this method, different statistical calculations could be performed to determine the metabolites’ capacities that allow the defining changes of a specific compound in definite situations [33]. The number of primary and secondary metabolites in a single organism may range from several hundred to tens of thousands, with little variation across orders of magnitude in concentration. Some strategies developed for metabolites analysis include metabolic profiling, metabolic fingerprinting, and target analysis [34].
Metabolic profiling is expected a simultaneous measure of a set of metabolites in a sample. Several analytical techniques can be used for metabolic profiling, such as (GC-MS), (LC-MS), and (NMR). To date, GC-MS is the most advanced analytical approach to metabolic profiling in plants [35]. Using GC-MS, it is possible to recognize several hundred compounds belonging to various classes, including sugar, organic acids, amino acids, alcohols, amines, and fatty acids. Similarly, LC-MS provides a better alternative for non-volatile compounds. The importance of LC-MS is increasing in metabolomics, especially after the adoption of ultra-performance liquid chromatography technology that can increase separation efficiency and decrease analysis time [36]. Substantially, NMR spectroscopy offers an entirely different analytical technique compared to MS-based approaches. The sensitivity of the NMR technique is much lower than MS-based techniques; however, the structural content information, reproducibility, and computable aspect could be superior to them [37]. Moreover, the preparation of the sample is simple, more convenient, and non-destructive measurement may possible. These properties of NMR make it an ideal tool for the identification of metabolites through metabolic profiling [38].

3. Workflow of Plant Metabolomics Analysis

The metabolomics of plants is very complex and varied in their chemical structure. Extensive identifications and a wide range of metabolic depictions could be attained with the arrangement of two or more metabolomics approaches and analytical methods, with the difference in extraction protocols [39]. Metabolomics analyses comprise the following three key tentative methods: (1) sample preparation, (2) data gaining, and (3) the identification of compounds by using the statistical analysis of the data. The preparation of the sample is a key step because it can contribute to the identification of a wide range of metabolites, which is comprised of tissue collecting, drying, or quenching, and metabolite extraction for analysis (derivatization) [40]. Thus, care should be taken in this step to avoid engaging in undesirable variation that can significantly disturb the analysis results. Many methods of enzyme quenching, such as drying, enzyme inhibitors and acids, and high meditations of organic solvents, could also distress the analysis and identification [41].
Plant metabolites are structurally different with high complications, such as dissimilar size, solubility, explosive nature, separation, amount, and stability [42,43]. The extraction method of metabolites relies on varied factors such as the type of plant organs, physical and chemical properties of the targeted metabolites, chemical structure, and the solvent used [44]. Generally, metabolite extraction methods include solvent extraction, supercritical fluid extraction, solid-phase extraction, and sonication [45]. Moreover, other methods are used to extract the essential oils, such as hydrodistillation, vapor-hydrodistillation, vapor-distillation, hydro diffusion, organic solvent extraction, and cold pressing [46]. Though, it is critically essential to evaluate metabolite extraction methodologies for a precise metabolite extraction study because a solvent composition that is good for one chemical class may not be suitable for another chemical class. Moreover, this could not be appropriate for extracting large numbers of metabolites from a specific tissue. So, it is important to understand and monitor the effects of the applied solvent treatment on the sample's metabolic content and profile obtained [47].
The measurement of complex metabolites needs an advanced analytical platform for sample analysis. Every platform’s range has a particular constraint, maybe in selectivity or sensitivity [48]. The selection of the analytical platform relies on the study initiated, the group of compounds, and their physiochemical properties, such as polarity, solubility, volatility, and concentration levels [49]. Additionally, one issue is that metabolites occur in a wide dynamic range of concentrations such as nanomolar and millimolar in the plant body. Subsequently, another problem is that not every metabolite is present in each tissue [50].
However, the most applied metabolomics approaches in analytical studies are liquid or gas chromatography synchronized with mass spectrometry (LC/GC-MS) and nuclear magnetic resonance spectrometry (NMR) [51]. Subsequently, another report [52] demonstrated an integrated technique that combines metabolites extraction and analysis with proteomic and RNA from a single sample that permits the immediate inquiry of all molecular levels and examines their interrelation and co-variance structure [53]. Consequently, biochemical regulation could result in the co-variance design of molecular dynamics in a cellular system [54]. In the context of metabolomics, the block diagram (Figure 1) of a typical experiment shows the following key steps:
  • Sample collection and organization;
  • Metabolites extraction;
  • Derivatization and separation;
  • Data acquisition;
  • Data analysis;
  • Metabolites identification;
  • Data submission to public repositories.

4. Metabolomics for Plant Stress Responses

Metabolomics is the scientific study of the set of metabolites present within an organism, plant cell, or tissue [55]. However, plant stress is any amendment in the growth and developmental conditions that distracts metabolic homeostasis and needs to modify the metabolic pathways in a process generally designated as acclimatization [35]. Over the last decade, metabolomics has developed promptly and is recognized as the prevailing technology in changing climatic conditions and assessing or elucidating testing phenotypes in assorted living systems [56]. Substantially, it may contribute to studying stress biology in plants or other organisms by recognizing various molecules, such as by-products of stress metabolism and compounds of stress signal transduction and related to the plant acclimation responses [52]. Their application has been driven in several fields, including medicinal, imitation biology, or analytical molding of plants, animals, and microscopic organisms [57].
Additionally, to the applicability of other fields, nowadays metabolomics could also be used on a large scale in the assortment procedure of plants and resistant to the varying environmental states. Different findings revealed that drought stress, salinity, extreme temperature, and soil flooding could cause significant instabilities in the pattern of plant metabolome [22]. Metabolomics signifies the ultimate omic’s level in a living system or reveals modifications in the traits of an organism or function. Different findings show the study of metabolomics under several environmental abiotic stresses, such as temperature [58], salinity and drought [59], and soil flooding [60]. In the same way, various metals and metalloids including, sulfur [61], phosphorus [62], oxidative stress [63], TM [64], and the combination of other several stress factors [65] in plants (Table 1). Various environmental factors that could negatively disturb the homeostasis and growth of plants are shown below (Figure 2).

4.1. The Response of Primary Metabolites to Abiotic Stresses

Plants established several adaptive mechanisms to endure abiotic factors, containing variations of metabolism in various directions, to confirm their existence in combative environmental situations [66] (Table 2). Several plant metabolites could assist and reduce the effect of the harsh stress of salt, drought, and water by acting as osmolytes and osmoprotectants [67]. Examples of such metabolites include dimethylsulfoniopropionate (DMSP) and glycine betaine; sugars, such as sucrose, trehalose, and fructan; amino acids, such as proline and ectoine, as well as some metabolites of polyols, sorbitol, and mannitol [68,69]. In plants, a wide range of waxy layers known as epicuticular wax keeps water balance during water shortage and acts as a mechanical stoppage to encounter disease-causing agents. Additionally, ascorbic acids, glutamine, alpha-tocopherol, anthocyanins, and carotene shield plant tissues by foraging the intermediates of bustling oxygen produced during oxidative stress [70]. Similarly, several other smaller compounds guard plants against oxidation damage related to various constrictions [65].
Besides, the plant’s defense system is related to generating phytoalexins, stimulating the common phenylpropanoid pathway and producing lignin biosynthesis [71]. Further, phytochemicals and hormones such as salicylic acid and methyl salicylate, methyl jasmonate and jasmonic acid, as well as other small molecules formed due to stress, play a significant role against environmental stresses [72,73,74]. All of these may also function as signaling compounds by stimulating the resistance system and reactions of acclimation [75]. Among the defense systems of plants, osmotic regulation is one of the broadly pronounced responses to the water shortage that needs the accretion of harmonious solutes, such as sugars, amino acids, polyols, and glycine betaine [76]. These chemical compounds do a significant job in sustaining cell turgor and stabilizing cell membranes and protein. Moreover, other studies designate the importance of these compounds in rehabilitating redox stability through the scavenging of ROS, which could adversely affect cellular structures and metabolism [68,77].
Table 1. List of species, various metabolomics approaches, and applications cited in this review under diverse abiotic stresses.
Table 1. List of species, various metabolomics approaches, and applications cited in this review under diverse abiotic stresses.
SpeciesAbiotic FactorsMethodApplicationReferences
Arabidopsis thaliana L.TemperatureGC-MSExploring the temperatures stress metabolome[58,78]
Populus euphratica Oliv.Water and salinityMetabolite profilingChanges in early and late transcription and metabolite profiles[79,80,81,82,83]
Thellungiella halophila (C.A.Mey.) O.E.SchulzMetabolic fingerprintingIdentify metabolic
changes in fruits
Solanum lycopersicum L.Metabolic fingerprintingTo classify control as well as salt-treated groups of tomatoes
Arabidopsis thaliana L.GC/MS and
LC/MS
To reveal the
short-term responses to salt stress
Arabidopsis thaliana L.Drought and
flooding
Metabolic profilingWhen defense pathways collide[22]
To identify the responses of plants to abiotic stresses[84]
Arabidopsis thaliana L.SulfurMulti-parallel, high-throughput analysisTo reveal novel findings[85]
Phaseolus vulgaris L.PhosphorusTranscript profilingTo investigate global gene expression and metabolic responses[86]
Arabidopsis thaliana LOxidativeGC-MSTo characterize the dynamics of metabolic[87]
Arabidopsis thaliana LHeavy metals
Caesium (Cs)
Cadmium (Cd)
NMRChange metabolic consequences of stress[88]
Silene cucubalus WibelMetabolomics analysis of the consequences of cadmium exposure[89]
Glycine max L.SalinityGC-MSMetabolomics analysis in the roots of different soya been varieties, under salinity levels[90]
Glycine. max L.CE-MSCE-MSProteomic profile investigation of different soya bean varieties, under Cd stressed conditions[91]

4.1.1. Amino Acids

Amino acids are considered a precursor for protein and other organic molecules, e.g., nucleic acids, which designate an active part in the responses of a plant under several stress factors. Amino acids could also play a significant role in signaling and controlling molecules [92]. Various studies showed that many amino acids stored in plants are apparent to different abiotic stresses [93,94]. Moreover, the exposure of plants to such stresses appearance an accumulation of proline and other amino acids. In plants, the role played by stored amino acids differs after acting as an osmolyte to adjust ions passage, reducing stomatal opening and reclamation of TM [95]. Moreover, amino acids can also disturb the synthesis and activity of several enzymes, gene expression, and redox state of homeostasis [96]. The accumulation of proline and ectoine is considered the most extensively dispersed osmolytes, as they act as osmoprotectants to protect plants from harmful effects and exciting environmental stresses, including low and high temperature, salinity, UVr, water, and osmotic stresses [68,97].
Primarily, proline is produced from a glutamate and proline metabolizing enzyme, pyrroline-5-carboxylate synthetase (P5CS), which reduces glutamate to pyrroline-5-carboxylate (P5C). At last, from the reduction of P5C, this stress-responsive amino acid forms by pyrroline-5-carboxylate reductases (P5CR) [98]. In transgenic plants, the significant role of proline was established during osmotic stress. For example, overexpression of the P5CS gene in soybean increased proline content and, thus, tolerance to salt stress in transgenic plants [99]. Besides osmolytes, proline is thought to accomplish many other important functions related to plant resistance, e.g., ROS scavenging, redox balancing, cytosolic pH buffer, molecular chaperon, and a stabilizer of protein structure [98]. Subsequently, in response to abiotic factors, the enlarged levels of proline were observed for several years to be the stress-responsive feature in plants. The relationship between the accumulation of proline as osmolytes and stress tolerance had a great share because of its applicability to different crops [100,101].
Remarkably, some of the metabolites were related to drought resistance and drought vulnerability of the considered hybrids [102]. Additionally, studies on drought responses at metabolomics levels indicated that Andean potatoes with a phenotype designating greater stress exposure have more proline related to the genetically assembled plant that was a higher dearth-tolerant [103]. It was established that the cultivar with a sensitive phenotype has high-level certain amino acids, containing proline and Gamma-aminobutyric acid (GABA) when barley exposed to salinity stress [104]. It may well advocate a greater liability of these plants to such stress. According to [96], this accretion could be associated with the deterioration of the leaf and slowing the development of a more subtle genotype. Furthermore, studies on Arabidopsis revealed that proline could be a lethal compound under heat stress [105], while Charlton et al. found that water deficiency was the cause of the decrease in isoleucine concentration in Pea and Arabidopsis plants [106].

4.1.2. Polyamines

Plants are tested by different stress factors and adversely affect their growth, yield, and geographical circulation [107]. To survive the combative environmental stress circumstances, plants have developed many adaptive strategies, amongst which the accumulation of metabolites plays an important defensive role [108]. Metabolites strongly involved in stress resistance are the low-molecular-weight (LMW) acyclic polyamines [109]. Polyamines are the LMW nitrogen-containing organic compounds with more than two amino groups with a positive charge at the cellular pH, allowing them to link with negatively charged molecules, such as nucleic acids, phospholipids, and proteins [110]. Usually, polyamines are polycations essential for plant growth and development and play an important role in abiotic stress resistance in higher plants. Triamine spermidine, tetraamine spermine, and their diamine predecessor, putrescine, are the general polyamines [111]. Because of their cationic nature, these compounds have often been correlated to environmental stresses, such as drought, chilling, heat, TM, and salinity [112].
The results of Khan et al. [95] and Capell et al. [113] showed that the accumulation of spermidine with the up-regulation of spermidine synthase of Cucurbita ficifolia augmented several stress responses in a recombinant Arabidopsis plant, such as waterlogging and salinity stresses. It was shown that spermidine acts as a signaling molecule and controls the assertion of intricate genes in drought resistance. Furthermore, it has been demonstrated that polyamines are attributed to being involved in maintaining membranes shielding from damage under stressful environments [114] and controlling the formation of nucleic acid as well as enzyme activity [115]. Additionally, different findings revealed that polyamines play a significant role in oxidative stress by mitigating the balance state of ROS through their direct contact or indirectly regulating the antioxidant system and suppressing ROS production. Moreover, some authors hypothesized that polyamines could act as a cellular signal in plants throughout the stress responses [116].

4.1.3. Carbohydrates

Carbohydrates produced during photosynthesis are the main building units that provide energy and support to the plant biomass [117]. Extensive studies revealed that non-living factors lead to the assemblage of non-structural saccharides, such as sucrose and lactose, simple sugars, or polyhydric compounds (alcohols and phenols), amongst various species of plants [118]. Particularly, there is a robust association between carbohydrate accretion and osmotic stress resistance, including oxidative stress (ROS) conditions, salt stress, and the scarcity of water [95]. As a source of carbon and energy in a cell, soluble carbohydrates may take a significant part in the metabolic processes of plants. Several stress factors may impact the level of these soluble carbohydrates because the accumulation of carbohydrates is associated with photosynthesis [119]. Rosa et al. [120] demonstrated that certain soluble sugars, such as sucrose and hexoses, improved stress tolerance by down-regulating the stress-related genes and up-regulating growth-related genes. Though, the contents of certain carbohydrates, such as raffinose, glucose, fructose, and maltose, are highly sensitive to environmental stresses and increase. However, the contents of myoinositol were reduced in barley roots during water-scarce conditions [121]. The findings of Sperdouli and Moustakas [122] revealed an increase and contact of augmented soluble carbohydrates, sustaining a great antioxidant defense in the leaves of Arabidopsis thaliana under dry environmental stress conditions. Studies showed renovation of carbon metabolism under salt-related stress (paraquat) in A. thaliana tissues and inferred by the researchers as a substitute approach to staying alive [122].
In water-deficit conditions, soluble sugars function as osmoprotectants, decreasing the harmful impact of osmotic stress and helps in sustaining the turgidity of cell and cell membrane stability by keeping plants from humiliation [123]. Under stress conditions, the increase in sugar quantity is generally the result of carbohydrate hydrolysis that needs enzymes with hydrolytic usage [124]. Moreover, carbohydrates that are soluble, such as disaccharides (sucrose and trehalose), oligosaccharides (raffinose and stachyose), and polymer of fructose molecules (fructans) next to their linked metabolic enzymes are essential compatible osmolytes associated with the scavenging of unstable molecules (ROS) during their assortment in plant tissues [125]. In low-temperature stress, sugar alcohols, such as polyols, function as osmoprotectants and shield cell membranes against ice adhesion [77]. Moreover, carbohydrates may act as signaling molecules [126]. The demonstrated data advocate a specific response of carbohydrates in plants. However, it should be noted that the accumulation of carbohydrates depends on the kind of stress to which it bared [127].

4.1.4. Glycine Betaine

Glycine betaine (GB) is a widely studied quat compound, which is active in retaining the water balance between the plant cell and the environment during drought conditions. Moreover, GB playing a significant role in stabilizing the macromolecules, shielding photosynthesis, detoxification of reactive oxygen radicals, and as an osmoprotectant [128,129]. Several studies indicated their importance in improving plant tolerance under various abiotic factors. It has been shown that plants are distinguished according to the formation of GB, such as barley, spinach, maize, and wheat, produce and accumulate a higher quantity of GB in their chloroplast. However, some plant species cannot obtain substantial amounts of GB during stress, such as A. thaliana, rice, and tobacco [130]. Furthermore, it has been shown that transgenic plants could mitigate the impact of abiotic stresses. Therefore, efforts have been made to improve tolerance through glycine betaine biosynthesis to achieve transgenic plants. In transgenic plants, such as Arabidopsis, the cyanobacteria genes, such as glycine sarcosine methyltransferase, and in transgenic maize, a greater amount of GB accumulates. As a result, in transgenic Arabidopsis, resistance to drought and salt is greater; nevertheless, a recombinant plant of maize retained well in cold-related to non-transgenic cultivars [131,132].
Moreover, through genetic engineering, other transgenic plants with a GB-producing capacity have been achieved, including Brassica juncea and tobacco with greater tolerance to salt and chilling, indicating a progressive ability to propagate and grow well related to wild-type in abiotic environmental conditions [133,134]. Besides, transgenic tomatoes with GB synthesis were more resistant to cold stress and produced fruit at a rate from 10 to 30% higher than the wild type. [135]. Though, the meditations of GB produced in every transgenic plant were scarce to control the osmotic stress to which plants were exposed. Similarly, previous studies showed that GB could enhance root growth and reduce oxidative stress. Additionally, the exogenous application of GB improves the stress tolerance of Cr in chickpea plants [136] and salinity stress in wheat [137]. Consequently, further protecting approaches of GB, such as defense against ROS and heavy metals stress, should be considered, which may enhance the tolerance level [138].

4.1.5. Lipids

Lipids are a fundamental component of biological membranes, particularly the plasma membrane, which serves as the contact between the cell and its surroundings [72]. Lipids can be grouped into eight major types based on the chemical structure in conjunction with distinctive hydrophobic and hydrophilic components, such as fatty acids, glycerides, phosphoglycerides, sphingolipids, steroids, isoprenoids, glycolipids, and polyketides [139]. Being sessile organisms, plants are subjected to a wide variety of biotic and abiotic factors, such as temperature, drought, heavy metals, salinity, and pathogen attack. However, lipid-mediated signaling occurs in response to all these stressors (Figure 2). The plasma membrane, which is typically the signaling source of lipids, is commonly used by plants to sense these stimuli and transform the signal into subsequent biochemical metabolism. Generally, these are acclimating enzymes that have all been proposed as signaling lipids, such as phospholipases, lipid kinases, and phosphatases [140]. Commonly, lysophospholipid, fatty acid, phosphatides, triacylglycerol, inositol phosphate, oxylipins, sphingolipids, and nacylethanolamine are considered the major contributing signaling lipids molecules [141]. The conformation and activity of cellular proteins and metabolites are influenced by signaling lipids because they have the ability to temporarily attract molecular markers to the membrane.
The enzyme phospholipase A (PLA) is very important in the formation of fatty acids and lysophospholipids. Usually, lysophospholipids are present in very limited amounts in plant tissues; however, in stressed conditions such as freezing their quantity increases [142]. Some reports revealed the physiological role of lysophospholipids against various environmental stresses. Similarly, the phospholipase A2 (PlA2) has been shown to increase the production of some elicitors in poppy plants [143], while lysophosphatidyl-choline and lysophosphatidyl-ethanolamine act as signals transducers in arbuscular symbiosis in potato [144].
Fatty acids have also been demonstrated as stress-responsive lipids in plants. Oleic acids modulate nitric oxide-related proteins, thereby regulating nitric oxide and mitigating tolerance in Arabidopsis [145]. Moreover, fatty acids also regulate drought, salt, and heavy metals tolerance, as well as the wound-induced responses of pathogens/herbivores in plants [146]. Likewise, the responsive role of phosphatidic acid (PA), inositol polyphosphates, oxylipins, sphingolipids, and some other lipids have been studied in various plant species [147,148,149]. Some of the environmental stress factors under which the plant lipid responses were reported to include chilling, freezing, and wounding [150], pathogens [151], low-temperature stress [152], salt stress [153], and water and drought [154] stress response.

4.2. The Response of Secondary Metabolites to Abiotic Stresses

Primary metabolites are compounds that are related to important physiological functions in organisms. Hence, they are generally found in all plant species and are directly involved in growth, development, and reproduction [155]. Compared to primary metabolites, secondary metabolites are very definite in their function, as they are not directly involved in plant growth, development, and reproduction of organisms. Generally, they are species-specific that could be redundant in different situations [156]. Usually, they are made under particular conditions for a definite purpose, such as defense against pathogens infection, enhanced resistance to abiotic stresses, and protect the harmful effect of UVr [157]. Furthermore, secondary metabolites produce different compounds important for several biochemical and biophysical processes in plant cells and tissues (Table 2). However, they have no common familiar physiological functions in plants, such as photosynthesis, respiration, translocation, transportation of solute, acclimatization of nutrients, and differentiation [158].
In addition, the specified plant species produce these natural products, and their concentration level is controlled to some extent with the growing period, environments, and adjustment progress [159]. Substantially, they attracted insects and animals for fertilization and seed spreading. The accumulation of phenyl amides in beans to the impact of abiotic factor (heat) was described, proposing an antioxidant role of these secondary metabolites [160]. Modern research tries to identify the key roles the secondary metabolites play in plants as indicators, antioxidants, and for other purposes. Secondary metabolites are also important in plants used by humans [161]. Besides, the compounds of secondary plant metabolites are distinctive means of food essences, medicines, flavorings, and other industrial materials [162]. In plants, the accretion of certain metabolites frequently occurs exposed to different stress factors, such as several phytohormones, elicitors, TM, and signal transduction compounds [163,164,165].
Some famous examples of secondary plant metabolites with medicinal properties include the anesthetic and antipyretic compounds salicin taken from Salix sp., which is used to make aspirin [166]. Similarly, other pharmacological secondary metabolites, such as taxol (anticancer), sequestered from pacific yew (Taxus brevifolia), and the strong obsessive compound morphine removed from opium (Papaver somniferum). Secondary metabolites have the following three major groups: phenolics, terpenes, and S and N comprising compounds (Figure 3) [167,168].

4.2.1. Phenolic Compounds

In plants, phenolic compounds are recognized as the largest and essential group of secondary metabolites changing from simpler aromatic rings to more complicated ones, such as lignin, and play a significant physiological role in increasing the resistance and adaptableness suboptimal circumstances during the life cycle of plants [169,170]. Phenolics are produced in optimum and sub-optimum environments in plants and play a major role in various developmental mechanisms, such as cell division, balancing hormones, photosynthetic processes, and reproduction, as well as in the mineralization of nutrients [75]. These compounds constitute secondary metabolites, including lignins and tannins, flavonoids, isoflavonoids, anthocyanins, and coumarins [171]. Moreover, all these chemical compounds are produced in plants by the phenylpropanoid pathway, in that phenylalanine compound is the main substratum that can do significant work in the resistance mechanism of plants against various stress factors in the environment [172].
The pathway of phenylpropanoid is regulated by biotic and abiotic factors, including drought, salt stress, TM, low or high temperature, wounding, pathogen attack, herbicide treatment, nutrient deficiencies, and UV radiations causing the accumulation of different phenolic compounds [75,173]. Consequently, the aggregations of phenolic compounds in plant materials are considered an important sustaining strategy of plants in harsh environmental situations. Hence, respond to these stresses and contribute to the removal of ROS, catalyzed-oxygenated reaction with the establishment of metabolic structures and obstructing the processes of oxidative enzymes, thus increasing evolutionary aptness [174,175]. Besides, phenolic accumulation is also considered a reliable feature and key defense mechanism under stress, leading to the enhanced creation of free radicals and other oxidative species in plants [176].
Moreover, to survive in oxidative stress conditions, plants had established two diverse biological ways, such as escaping ROS creation and eliminating it through enzymatic and non-enzymatic processes, such as the deposition of LMW antioxidants [177,178]. Further, studies revealed that the accumulation of LMW antioxidants results due to the activities of phenylalanine ammonia-lyase (PAL), chalcone synthase (CHS), and other essential enzymes [179]. Various physiological processes of plants related to growth and development in plants comprising seed germination, cell division, and synthesis of photosynthetic pigments, are influenced by phenolic acids and flavonoid accumulation for persistence and adaptation to environmental conflicts [180,181]. In particular, phenolic compounds consult greater tolerance in plants such as TM stress, which enhances the production of ROS and reduced growth [182], and phenolic compounds (flavonoids), in response, protect plants from oxidative stress damage through the chelation process [183,184]. Similarly, when plants are exposed to other abiotic factors can also affect their life cycle. Under drought conditions, the concentration of ferulic acid decreased, while the p-coumaric acid and caffeic acid increased in maize xylem sap, which could be supportive in stiffening and lignification of the cell wall [185]. Spatially confined fluctuations in cell wall phenolics were presented to be engaged in the advanced inhibition of wall extensibility and root growth, which can enable root acclimation to drought [186].
Various environmental stress factors mediated the synthesis of flavonoids, isoflavonoids, and anthocyanins. In plants, flavonoids play a defensive role due to their antioxidant properties when exposed to a water-deficit situation [187]. Moreover, Nakabayashi et al. [188] indicated that flavonoid significantly improves resistance in A. thaliana in water scare conditions. Similarly, phenolic acids and flavonoids as antioxidants and sunshades are involved in plants’ response to a dry environment [189]. Other studies reported that different polyphenolic was associated with gene expression reforms in an account of potato plants under drought conditions, though the fluctuations were greatly specific to the cultivar [190]. Rodziewicz et al. [96] and Parida et al. [191] suggested that polyphenols are involved in conserving osmotic potential in cells and confiscating free radicals during drought stress. Besides, polyphenols affect the source and movement of organic and inorganic soil nutrients existing for plants and microbes and indicate a reply to nutrient insufficiency, therefore offering a way for identifying nutrient ailments earlier to the occurrence of evident symptoms [192]. Stress conditions of drought and waterlogging increased the flavonoids quercetin and rutin in the herbaceous pharmaceutical plant Hypericum brasiliense, whereas cold stress caused a different reaction [193]. Comparatively, a greater decrease in flavonoids was noticed in the sensitive genotypes; thus, they could show that the flavonoid content was imperative in sustaining the greater antioxidant activity in water-stressed conditions [96,194].
Furthermore, anthocyanins were identified to increase their content in plant tissues against drought and cold stress because of their antioxidant and ROS scavenging properties, which cause protection to plant cells [195,196]. In red-fleshed apple callus culture, low temperature (16 °C) tempted an increased level of anthocyanin [197].
Additionally, the assembly of phenolics rises into the cell wall either as suberin or lignin under low-temperature stress [198]. Though, suberin deposition and lignification increase the adaptability and resistance to cold stress [192]. Similarly, to respond to the negative effects of Uvr, endogenous phenolic compounds (flavonoids) accumulate in plant cells and make a shield under the epidermal layer, which protects the plant and the component of the cell from these harmful radiations [199]. Moreover, flavonolignan silymarin has been reported to accumulate in in-vitro cultures of Silybum marianum upon application of abiotic stress treatments, such as NaCl, polyethylene glycol, and gamma irradiation, because of the defensive mechanisms that cells perform to counteract the stress of these factors [200]. Therefore, stress elicitation successfully produces high-value phytoconstituents from medicinal crops [201,202].

4.2.2. Terpenoids

Plants have developed a complex resistance system that depends on the swift perception and instigation of secondary metabolites to adopt different environmental stress factors in an ecosystem [203]. Among all, terpenoids establish a broad and structurally diverse group of lipophilic secondary metabolites, which are produced in plants from isoprene units (C5H8) [204]. Physiologically, terpenoids play an important role as phytohormones, such as the sesquiterpenoid abscisic acid (ABA) and the diterpenoids gibberellic acid (GA), against biotic and abiotic stresses. Studies have shown that the phytohormone abscisic acid (ABA) triggers defense mechanisms, such as facilitating responses to drought and water stress by adapting the membrane properties [205]. Moreover, terpenoids show antioxidant and antibiotic activity that maintains lipid membranes and increases environmental stress tolerance against herbivores [206].
Terpenoids also function as phytoalexins (LMW antimicrobial compounds), prepared as part of the plant defense mechanism in response to abiotic and biotic factors. For example, many diterpene phytoalexins have been reported in Oryza sativa [207]. Similarly, in cotton plants, sesquiterpenoid phytoalexins, such as gossypol, hemigossypolone, and heliocides, as defensive metabolites accumulated both above and below the ground against pathogens and herbivores [208]. Moreover, in maize and rice leaves and roots, diterpene phytoalexins are produced, including zealexins, kauralexins, and oryzalexins that exhibit antimicrobial properties and respond against pathogenic fungal blast diseases, such as rice blast caused by Magnaporthe grisea [207,209]. Additionally, UVr and TM stress induced the accumulation of rice phytoalexins. According to Vaughan et al. [209], the accumulation of phytoalexins in response to drought is root-specific and does not affect the level of phytoalexins aboveground. However, the reduced content of the terpenoid compound was described in cotton species in drought conditions [191,210]. Yusuf et al. [211] noticed that the increased content of soluble alcohol tocopherol with antioxidant properties shows a significant role in the mitigation of stress by stabilizing the cell membranes induced by salinity, TM, and osmotic potential in B. juncea. Furthermore, the content of saponins in soybean plants was recognized as one of the crucial secondary metabolites related to the resistance of salt stress [212].
Table 2. The response of various types of metabolites against different abiotic stresses.
Table 2. The response of various types of metabolites against different abiotic stresses.
MetabolomicsStressMode of ActionReferences
Primary Metabolites
Amino acids: (proline)Drought, salinity, temperature, and coldActs as osmoprotectant[77,96,97]
Polyamines:
(triamine spermidine, tetraamine, spermine)
Heavy metalsRegulating antioxidant systems, suppressing ROS production[168]
Carbohydrates:
a. (sugar, sucrose)
Water deficitOsmoprotectant, maintain turgor, cell membranes stability[95]
b. alcohols (sorbitol, ribitol, and inositol)Cold stressCryoprotectants protect cell membranes against ice adhesion[97,124]
c. disaccharides, raffinoseROSROS scavengers,
control ROS signaling
[39,106]
Glycine betaineDrought, ROS, salt, and low temperatureOsmoprotectant
detoxification of ROS,
[128]
LipidsHeavy metals stressScavenge the ROS production[136,138]
Secondary Metabolites
Phenolic compounds: p-coumaric acid, caffeic acid; flavonoids, anthocyanin, suberin, or ligninHeavy metals/ ROSScavenging of ROS and chelation process[151,159]
Water stressAntioxidant mechanism[154]
Drought, UVStiffening and lignification of the cell wall, antioxidant, and sun shields properties[191]
Drought, nutrient deficiencyScavenging of ROS, maintenance of osmotic potential in cells, and identifying nutrient ailments[77]
Cold and droughtIncrease resistance and protect plant cell[161]
Cold or low temperatureLignification and submarine deposition increase adaptability and resistance[213]
TerpenoidsAbscisic acid (ABA), gibberellic acid (GA), phytoalexins (gossypol, hemigossypolone and heliocides), momilactones, oryzalexins, tocopherol, saponinsBiotic and abiotic factorsPhysiological function, ameliorate heavy metal stresses, antioxidant, and antibiotic activity[170,171]
Heavy metal, drought, UV, pathogens, and herbivoresImprove stress tolerance, drought, heavy metals, and enhances antimicrobial properties[207,208,209,210]
Fungal blastStabilizing the cell membranes[212]
Salinity, heavy metal, potential osmoticSalt stress tolerance[212]
Nitrogen-containing metabolites
Alkaloids
Glucosinolates
Non-protein amino acids
Drought, herbivoresIncrease tolerance level and defense against herbivore attack[158,214]
Drought, waterloggingOsmoprotectants increased phytochemical contents[215,216]

4.2.3. Nitrogen-Containing Secondary Metabolites

Plants have developed several defense mechanisms against invading enemies, such as microbial pathogens and herbivorous animals, as well as abiotic factors, e.g., drought, waterlogging, and salinity, which are considered for the high loss of crop production worldwide [217]. However, plants have developed a complex defense system of secondary metabolism against these stressors, including the nitrogen-containing secondary metabolites, such as alkaloids, cyanogenic glycosides or glucosinolates, and non-protein amino acids (Figure 4) [158,217]. Previously, nitrogen-containing secondary metabolites were considered unwanted materials of plants and are known now for their resistivity towards different stress factors [119]. Among the phytochemicals, alkaloids are heterogeneous groups of secondary metabolites consisting of one or more nitrogen atoms produced under abiotic stress conditions. It has been found that alkaloids perform a significant role against microbial pathogens and herbivorous animals. Besides, more alkaloid contents and derivatives are produced in abiotic stress conditions. For example, poppy plants make more alkaloids when there is a drought period as well as under salinity stress [214]. In lupins (Lupinus termis) cultivars, the content of alkaloids was also influenced by the drought and activated yeast extract treatment [216].
Additionally, glucosinolates and cyanogenic glycosides are sulfur and nitrogen-containing secondary metabolites derived from glucose and amino acids. Similarly, Rodziewicz et al. [96] demonstrated that all these natural compounds play a significant role against different environmental factors (biotic and abiotic). Mewis et al. [215] showed that in A. thaliana, under drought and water logging conditions tend to increase aliphatic compounds of glucosinolate and flavonoids. Moreover, in B. juncea, the increased level of glucosinolate was observed during the vegetative stage under water deficit conditions. In plants, apart from the essential 20 amino acids, there are more than 200 free plant cell amino acids that are not assimilated into proteins. These free amino acids are called non-protein amino acids. Their major function in plants is to respond to various environmental stresses [158,217].

5. Conclusion and Future Perspectives

The acceleration of climate change increases the severity of damage to crop productivity under environmental stress. Understanding the role that primary and secondary metabolites play during stress resistance mechanisms is important for developing crop species and improving their stress resistance, ensuring that the need for food security is met for a growing global population. However, less has been understood about the function of these metabolites against environmental stresses in plants, especially abiotic stresses. In the current review, we have provided an overview of the role of primary (amino acids, polyamines, carbohydrates, glycine betaine, and lipids) and secondary (phenolics, flavonoids, terpenoids, alkaloids, and glycosides) metabolites against several abiotic factors, such as drought, salinity, temperature, UVr, and TM. Analysis of more than 200 articles allowed us to describe the main responses of primary and secondary metabolite products of different plant species to abiotic stresses. Metabolomics has occupied a prominent place in plant stress physiology and biology research. Metabolic change due to abiotic stress is complex to describe the variability between different plant species. Nevertheless, metabolomics needs more extensive research in data annotation, assessment, processing, and evaluation. Progress in “omics” tools and bioinformatics and enhanced assimilation of the data from varying molecular levels is needed. Hence, to expose the full picture of sustaining mechanism, which will lead to new biomarkers of resistance towards biotic and abiotic stresses. Affirmation of the impact of environmental stresses on plants and their metabolite level responses recorded valued genes about the mechanism underlying such acclimation. However, the balancing mechanism between the gene expression and the subsequent metabolic phenotype is a big challenge nowadays. Therefore, comprehensive research of the dynamic behavior of metabolic systems is a great task for researchers in systematic biology. Furthermore, identifying the genetic background behind the diversity of primary and secondary metabolites produced by plants will help in improving and developing stress tolerance. Manipulating and overexpressing genes related to the biosynthetic pathway of secondary metabolites could be a solution for plant tolerance to environmental stress conditions.

Author Contributions

Conceptualization: S.A. and U.S.; Investigation: Z.-H.T. and S.A.; Resources: S.U., U.S., A.A.E. and Y.K.; Software: S.U. and A.A.E.; Supervision: Z.-H.T.; Writing—original draft: S.U.; Review and editing: A.A.E., J.K., and Y.K.; Equal contribution: S.U.; References: A.K. All authors have read and agreed to the published version of the manuscript.

Funding

The key Research and Development project of Heilongjiang Province, China (JD22A008).

Acknowledgments

The authors are thankful to the China Scholarship Council (CSC) for support and to the Key Laboratory of Plant Ecology, Northeast Forestry University, Harbin 150040, China, for providing an excellent research environment. Moreover, Uzma Salam and Shakir Ullah are thankful to Zhong-Hua Tang for his continuous support and assistance during their studies.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Athar, H.; Ashraf, M. Strategies for crop improvement against salinity and drought stress: An overview. In Salinity and Water Stress; Springer: Berlin/Heidelberg, Germany, 2009; pp. 1–16. [Google Scholar]
  2. Khan, A.; Ali, S.; Khan, M.; Hamayun, M.; Moon, Y.-S. Parthenium hysterophorus’s Endophytes: The Second Layer of Defense against Biotic and Abiotic Stresses. Microorganisms 2022, 10, 2217. [Google Scholar] [CrossRef]
  3. Hayat, K.; Khan, J.; Khan, A.; Ullah, S.; Ali, S.; Fu, Y. Ameliorative effects of exogenous Proline on photosynthetic attributes, nutrients uptake, and oxidative stresses under cadmium in Pigeon pea (Cajanus cajan L.). Plants 2021, 10, 796. [Google Scholar] [CrossRef] [PubMed]
  4. Collier, P.; Dercon, S. African agriculture in 50 years: Smallholders in a rapidly changing world? World Dev. 2014, 63, 92–101. [Google Scholar] [CrossRef]
  5. Hayat, K.; Khan, A.; Bibi, F.; Murad, W.; Fu, Y.; Batiha, G.E.-S.; Alqarni, M.; Khan, A.; Al-Harrasi, A. Effect of Cadmium and Copper Exposure on Growth, Physio-Chemicals and Medicinal Properties of Cajanus cajan L.(Pigeon Pea). Metabolites 2021, 11, 769. [Google Scholar] [CrossRef]
  6. Ullah, S.; Khan, J.; Hayat, K.; Abdelfattah Elateeq, A.; Salam, U.; Yu, B.; Ma, Y.; Wang, H.; Tang, Z.-H. Comparative study of growth, cadmium accumulation and tolerance of three chickpea (Cicer arietinum L.) cultivars. Plants 2020, 9, 310. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Sung, J.; Lee, S.; Lee, Y.; Ha, S.; Song, B.; Kim, T.; Waters, B.M.; Krishnan, H.B. Metabolomic profiling from leaves and roots of tomato (Solanum lycopersicum L.) plants grown under nitrogen, phosphorus or potassium-deficient condition. Plant Sci. 2015, 241, 55–64. [Google Scholar] [CrossRef]
  8. Hasanuzzaman, M.; Nahar, K.; Alam, M.M.; Roychowdhury, R.; Fujita, M. Physiological, biochemical, and molecular mechanisms of heat stress tolerance in plants. Int. J. Mol. Sci. 2013, 14, 9643–9684. [Google Scholar] [CrossRef]
  9. Sun, C.; Liu, L.; Wang, L.; Li, B.; Jin, C.; Lin, X. Melatonin: A master regulator of plant development and stress responses. J. Integr. Plant Biol. 2021, 63, 126–145. [Google Scholar] [CrossRef] [PubMed]
  10. Zhang, A.; Sun, H.; Yan, G.; Wang, P.; Wang, X. Mass spectrometry-based metabolomics: Applications to biomarker and metabolic pathway research. Biomed. Chromatogr. 2016, 30, 7–12. [Google Scholar] [CrossRef]
  11. Hummel, J.; Strehmel, N.; Bölling, C.; Schmidt, S.; Walther, D.; Kopka, J. Mass spectral search and analysis using the golm metabolome database. In The Handbook of Plant Metabolomics; Wiley: Hoboken, NJ, USA, 2013; pp. 321–343. [Google Scholar]
  12. Salekdeh, G.H.; Komatsu, S. Crop proteomics: Aim at sustainable agriculture of tomorrow. Proteomics 2007, 7, 2976–2996. [Google Scholar] [CrossRef]
  13. Ali, I.; Khan, A.; Ali, A.; Ullah, Z.; Dai, D.-Q.; Khan, N.; Khan, A.; Al-Tawaha, A.R.; Sher, H. Iron and zinc micronutrients and soil inoculation of Trichoderma harzianum enhance wheat grain quality and yield. Front. Plant Sci. 2022, 13. [Google Scholar] [CrossRef]
  14. Ashraf, M.; Ahmad, M.S.A.; Öztürk, M.; Aksoy, A. Crop improvement through different means: Challenges and prospects. In Crop Production for Agricultural Improvement; Springer: Berlin/Heidelberg, Germany, 2012; pp. 1–15. [Google Scholar]
  15. Das, R.; Tzudir, L. Climate Change and Crop Stresses. Biot. Res. Today 2021, 3, 351–353. [Google Scholar]
  16. Minhas, P.S.; Rane, J.; Pasala, R.K. Abiotic stresses in agriculture: An overview. In Abiotic Stress Management for Resilient Agriculture; Springer: Berlin/Heidelberg, Germany, 2017; pp. 3–8. [Google Scholar]
  17. Coskun, D.; Britto, D.T.; Huynh, W.Q.; Kronzucker, H.J. The role of silicon in higher plants under salinity and drought stress. Front. Plant Sci. 2016, 7, 1072. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Alamri, S.; Hu, Y.; Mukherjee, S.; Aftab, T.; Fahad, S.; Raza, A.; Ahmad, M.; Siddiqui, M.H. Silicon-induced postponement of leaf senescence is accompanied by modulation of antioxidative defense and ion homeostasis in mustard (Brassica juncea) seedlings exposed to salinity and drought stress. Plant Physiol. Biochem. 2020, 157, 47–59. [Google Scholar] [CrossRef] [PubMed]
  19. Zlatev, Z.; Lidon, F.C. An overview on drought induced changes in plant growth, water relationsand photosynthesis. Emir. J. Food Agric. 2012, 24, 57–72. [Google Scholar]
  20. Khan, N.; Bano, A.; Rahman, M.A.; Rathinasabapathi, B.; Babar, M.A. UPLC-HRMS-based untargeted metabolic profiling reveals changes in chickpea (Cicer arietinum) metabolome following long-term drought stress. Plant Cell Environ. 2019, 42, 115–132. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Fernández, R.; Bertrand, A.; Reis, R.; Mourato, M.; Martins, L.; González, A. Growth and physiological responses to cadmium stress of two populations of Dittrichia viscosa (L.) Greuter. J. Hazard. Mater. 2013, 244, 555–562. [Google Scholar] [CrossRef]
  22. Obata, T.; Fernie, A.R. The use of metabolomics to dissect plant responses to abiotic stresses. Cell. Mol. Life Sci. 2012, 69, 3225–3243. [Google Scholar] [CrossRef] [Green Version]
  23. Cairns, J.E.; Prasanna, B.M. Developing and deploying climate-resilient maize varieties in the developing world. Curr. Opin. Plant Biol. 2018, 45, 226–230. [Google Scholar] [CrossRef]
  24. Begum, H.A.; Hamayun, M.; Khan, A.; Yaseem, T.; Bussmann, R.W.; Murad, W. Quantitative ethnobotanical appraisal of medicinal plants used by indigenous communities of District Malakand, Pakistan. Ethnobot. Res. Appl. 2022, 24, 1–14. [Google Scholar]
  25. Bulgari, R.; Franzoni, G.; Ferrante, A. Biostimulants application in horticultural crops under abiotic stress conditions. Agronomy 2019, 9, 306. [Google Scholar] [CrossRef] [Green Version]
  26. Ghag, S.B.; Ganapathi, T.; Jain, S.M.; Penna, S. Omics technologies and breeding of horticultural crops. In Omics in Horticultural Crops; Elsevier: Amsterdam, The Netherlands, 2022; pp. 75–90. [Google Scholar]
  27. Razzaq, A.; Sadia, B.; Raza, A.; Khalid Hameed, M.; Saleem, F. Metabolomics: A way forward for crop improvement. Metabolites 2019, 9, 303. [Google Scholar] [CrossRef] [Green Version]
  28. Piestansky, J.; Matuskova, M.; Cizmarova, I.; Majerova, P.; Kovac, A.; Mikus, P. Ultrasensitive determination of serotonin in human urine by a two dimensional capillary isotachophoresis-capillary zone electrophoresis hyphenated with tandem mass spectrometry. J. Chromatogr. A 2021, 1648, 462190. [Google Scholar] [CrossRef] [PubMed]
  29. Chen, D.; McCool, E.N.; Yang, Z.; Shen, X.; Lubeckyj, R.A.; Xu, T.; Wang, Q.; Sun, L. Recent advances (2019–2021) of capillary electrophoresis-mass spectrometry for multilevel proteomics. Mass Spectrom. Rev. 2021, 42, 617–642. [Google Scholar] [CrossRef]
  30. Rodrigues, A.M.; Ribeiro-Barros, A.I.; António, C. Experimental design and sample preparation in forest tree metabolomics. Metabolites 2019, 9, 285. [Google Scholar] [CrossRef] [Green Version]
  31. Duhoux, A.; Carrère, S.; Gouzy, J.; Bonin, L.; Délye, C. RNA-Seq analysis of rye-grass transcriptomic response to an herbicide inhibiting acetolactate-synthase identifies transcripts linked to non-target-site-based resistance. Plant Mol. Biol. 2015, 87, 473–487. [Google Scholar] [CrossRef] [PubMed]
  32. Fiehn, O.; Robertson, D.; Griffin, J.; van der Werf, M.; Nikolau, B.; Morrison, N.; Sumner, L.W.; Goodacre, R.; Hardy, N.W.; Taylor, C. The metabolomics standards initiative (MSI). Metabolomics 2007, 3, 175–178. [Google Scholar] [CrossRef] [Green Version]
  33. van den Berg, R.A.; Hoefsloot, H.C.; Westerhuis, J.A.; Smilde, A.K.; van der Werf, M.J. Centering, scaling, and transformations: Improving the biological information content of metabolomics data. BMC Genom. 2006, 7, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Gholkar, M.S.; Li, J.V.; Daswani, P.G.; Tetali, P.; Birdi, T.J. 1H nuclear magnetic resonance-based metabolite profiling of guava leaf extract: An attempt to develop a prototype for standardization of plant extracts. BMC Complement. Med. Ther. 2021, 21, 1–20. [Google Scholar] [CrossRef] [PubMed]
  35. Shulaev, V.; Cortes, D.; Miller, G.; Mittler, R. Metabolomics for plant stress response. Physiol. Plant. 2008, 132, 199–208. [Google Scholar] [CrossRef] [PubMed]
  36. Giri, S.; Krausz, K.W.; Idle, J.R.; Gonzalez, F.J. The metabolomics of (±)-arecoline 1-oxide in the mouse and its formation by human flavin-containing monooxygenases. Biochem. Pharmacol. 2007, 73, 561–573. [Google Scholar] [CrossRef] [PubMed]
  37. Schripsema, J. Application of NMR in plant metabolomics: Techniques, problems and prospects. Phytochem. Anal. Int. J. Plant Chem. Biochem. Tech. 2010, 21, 14–21. [Google Scholar] [CrossRef] [PubMed]
  38. Krishnan, P.; Kruger, N.; Ratcliffe, R. Metabolite fingerprinting and profiling in plants using NMR. J. Exp. Bot. 2005, 56, 255–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Pan, H.; Kou, P.; Yang, J.; Niu, L.; Wan, N.; Zhao, C.; Liu, Z.; Gu, C.; Fu, Y. A novel approach for efficient extraction and enrichment of phytochemicals with CO2-based switchable-solvent from pigeon pea leaves. J. Clean. Prod. 2021, 284, 124629. [Google Scholar] [CrossRef]
  40. Pan, H.; Nie, S.; Wang, Z.; Yu, L.; Liu, Z.; Xu, J.; Fu, Y. Ultrasound-assisted extraction of phytochemicals from Cili leaves with a novel CO2-responsive surfactant-aqueous and extraction mechanism. Ind. Crops Prod. 2022, 175, 114241. [Google Scholar] [CrossRef]
  41. Kim, H.K.; Verpoorte, R. Sample preparation for plant metabolomics. Phytochem. Anal. Int. J. Plant Chem. Biochem. Tech. 2010, 21, 4–13. [Google Scholar] [CrossRef]
  42. Halket, J.M.; Waterman, D.; Przyborowska, A.M.; Patel, R.K.; Fraser, P.D.; Bramley, P.M. Chemical derivatization and mass spectral libraries in metabolic profiling by GC/MS and LC/MS/MS. J. Exp. Bot. 2005, 56, 219–243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Twaij, B.M.; Hasan, M.N. Bioactive Secondary Metabolites from Plant Sources: Types, Synthesis, and Their Therapeutic Uses. Int. J. Plant Biol. 2022, 13, 4–14. [Google Scholar] [CrossRef]
  44. Pan, H.; Nie, S.; Kou, P.; Wang, L.; Wang, Z.; Zhao, C.; Wang, X.; Fu, Y. An enhanced extraction and enrichment phytochemicals from Rosa roxburghii Tratt leaves with ultrasound-assist CO2-based switchable-solvent and extraction mechanism study. J. Mol. Liq. 2021, 337, 116591. [Google Scholar] [CrossRef]
  45. Jin, W.; Zhou, T.; Li, G. Recent advances of modern sample preparation techniques for traditional Chinese medicines. J. Chromatogr. A 2019, 1606, 460377. [Google Scholar] [CrossRef]
  46. El Asbahani, A.; Miladi, K.; Badri, W.; Sala, M.; Addi, E.A.; Casabianca, H.; El Mousadik, A.; Hartmann, D.; Jilale, A.; Renaud, F. Essential oils: From extraction to encapsulation. Int. J. Pharm. 2015, 483, 220–243. [Google Scholar] [CrossRef]
  47. Mahmud, I.; Sternberg, S.; Williams, M.; Garrett, T.J. Comparison of global metabolite extraction strategies for soybeans using UHPLC-HRMS. Anal. Bioanal. Chem. 2017, 409, 6173–6180. [Google Scholar] [CrossRef] [PubMed]
  48. Ellis, D. Biomarker metabolites capturing the metabolite variance present in a rice plant developmental period. Physiol. Plant. 2008, 5, 8. [Google Scholar]
  49. Allwood, J.W.; Goodacre, R. An introduction to liquid chromatography–mass spectrometry instrumentation applied in plant metabolomic analyses. Phytochem. Anal. Int. J. Plant Chem. Biochem. Tech. 2010, 21, 33–47. [Google Scholar] [CrossRef] [PubMed]
  50. Wang, J.H.; Byun, J.; Pennathur, S. Analytical approaches to metabolomics and applications to systems biology. In Seminars in Nephrology; Elsevier: Amsterdam, The Netherlands, 2010. [Google Scholar]
  51. Wu, H.; Southam, A.D.; Hines, A.; Viant, M.R. High-throughput tissue extraction protocol for NMR-and MS-based metabolomics. Anal. Biochem. 2008, 372, 204–212. [Google Scholar] [CrossRef]
  52. Ghatak, A.; Chaturvedi, P.; Weckwerth, W. Metabolomics in plant stress physiology. In Plant Genetics and Molecular Biology; Springer: Berlin/Heidelberg, Germany, 2018; pp. 187–236. [Google Scholar]
  53. Valledor, L.; Escandón, M.; Meijón, M.; Nukarinen, E.; Cañal, M.J.; Weckwerth, W. A universal protocol for the combined isolation of metabolites, DNA, long RNA s, small RNA s, and proteins from plants and microorganisms. Plant J. 2014, 79, 173–180. [Google Scholar] [CrossRef]
  54. Wienkoop, S.; Morgenthal, K.; Wolschin, F.; Scholz, M.; Selbig, J.; Weckwerth, W. Integration of metabolomic and proteomic phenotypes: Analysis of data covariance dissects starch and RFO metabolism from low and high temperature compensation response in Arabidopsis thaliana. Mol. Cell. Proteom. 2008, 7, 1725–1736. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Guo, H.; Guo, H.; Zhang, L.; Tang, Z.; Yu, X.; Wu, J.; Zeng, F. Metabolome and transcriptome association analysis reveals dynamic regulation of purine metabolism and flavonoid synthesis in transdifferentiation during somatic embryogenesis in cotton. Int. J. Mol. Sci. 2019, 20, 2070. [Google Scholar] [CrossRef] [Green Version]
  56. Guy, C.; Kaplan, F.; Kopka, J.; Selbig, J.; Hincha, D.K. Metabolomics of temperature stress. Physiol. Plant. 2008, 132, 220–235. [Google Scholar] [CrossRef]
  57. Putri, S.P.; Nakayama, Y.; Matsuda, F.; Uchikata, T.; Kobayashi, S.; Matsubara, A.; Fukusaki, E. Current metabolomics: Practical applications. J. Biosci. Bioeng. 2013, 115, 579–589. [Google Scholar] [CrossRef]
  58. Cook, D.; Fowler, S.; Fiehn, O.; Thomashow, M.F. A prominent role for the CBF cold response pathway in configuring the low-temperature metabolome of Arabidopsis. Proc. Natl. Acad. Sci. USA 2004, 101, 15243–15248. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Javed, J.; Rauf, M.; Arif, M.; Hamayun, M.; Gul, H.; Ud-Din, A.; Ud-Din, J.; Sohail, M.; Rahman, M.M.; Lee, I.-J. Endophytic fungal consortia enhance basal drought-tolerance in Moringa oleifera by upregulating the antioxidant enzyme (APX) through Heat shock factors. Antioxidants 2022, 11, 1669. [Google Scholar] [CrossRef] [PubMed]
  60. Ackah, M.; Shi, Y.; Wu, M.; Wang, L.; Guo, P.; Guo, L.; Jin, X.; Li, S.; Zhang, Q.; Qiu, C. Metabolomics response to drought stress in Morus alba L. variety Yu-711. Plants 2021, 10, 1636. [Google Scholar] [CrossRef] [PubMed]
  61. Chung, J.-S.; Kim, H.-C.; Yun, S.-M.; Kim, H.-J.; Kim, C.-S.; Lee, J.-J. Metabolite analysis of lettuce in response to sulfur nutrition. Horticulturae 2022, 8, 734. [Google Scholar] [CrossRef]
  62. Li, Z.; Hu, J.; Wu, Y.; Wang, J.; Song, H.; Chai, M.; Cong, L.; Miao, F.; Ma, L.; Tang, W. Integrative analysis of the metabolome and transcriptome reveal the phosphate deficiency response pathways of alfalfa. Plant Physiol. Biochem. 2022, 170, 49–63. [Google Scholar] [CrossRef]
  63. Savchenko, T.; Tikhonov, K. Oxidative stress-induced alteration of plant central metabolism. Life 2021, 11, 304. [Google Scholar] [CrossRef]
  64. Feng, Z.; Ji, S.; Ping, J.; Cui, D. Recent advances in metabolomics for studying heavy metal stress in plants. TrAC Trends Anal. Chem. 2021, 143, 116402. [Google Scholar] [CrossRef]
  65. Patel, J.; Khandwal, D.; Choudhary, B.; Ardeshana, D.; Jha, R.K.; Tanna, B.; Yadav, S.; Mishra, A.; Varshney, R.K.; Siddique, K.H. Differential Physio-Biochemical and Metabolic Responses of Peanut (Arachis hypogaea L.) under Multiple Abiotic Stress Conditions. Int. J. Mol. Sci. 2022, 23, 660. [Google Scholar] [CrossRef]
  66. Ashapkin, V.V.; Kutueva, L.I.; Aleksandrushkina, N.I.; Vanyushin, B.F. Epigenetic mechanisms of plant adaptation to biotic and abiotic stresses. Int. J. Mol. Sci. 2020, 21, 7457. [Google Scholar] [CrossRef]
  67. Teklić, T.; Parađiković, N.; Špoljarević, M.; Zeljković, S.; Lončarić, Z.; Lisjak, M. Linking abiotic stress, plant metabolites, biostimulants and functional food. Ann. Appl. Biol. 2021, 178, 169–191. [Google Scholar] [CrossRef]
  68. Dikilitas, M.; Simsek, E.; Roychoudhury, A. Role of proline and glycine betaine in overcoming abiotic stresses. In Protective Chemical Agents in the Amelioration of Plant Abiotic Stress: Biochemical and Molecular Perspectives; Wiley: Hoboken, NJ, USA, 2020; pp. 1–23. [Google Scholar]
  69. Kebert, M.; Vuksanović, V. Species-Level Differences in Osmoprotectants and Antioxidants Contribute to Stress Tolerance of Quercus robur L., and Q. cerris L. Seedlings under Water Deficit and High Temperatures. Plants 2022, 11, 1744. [Google Scholar] [PubMed]
  70. Hüdig, M.; Schmitz, J.; Engqvist, M.; Maurino, V. Biochemical control systems for small molecule damage in plants. Plant Signal. Behav. 2018, 13, e1477906. [Google Scholar] [CrossRef] [PubMed]
  71. Kishi-Kaboshi, M.; Okada, K.; Kurimoto, L.; Murakami, S.; Umezawa, T.; Shibuya, N.; Yamane, H.; Miyao, A.; Takatsuji, H.; Takahashi, A. A rice fungal MAMP-responsive MAPK cascade regulates metabolic flow to antimicrobial metabolite synthesis. Plant J. 2010, 63, 599–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Hou, Q.; Ufer, G.; Bartels, D. Lipid signalling in plant responses to abiotic stress. Plant Cell Environ. 2016, 39, 1029–1048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Li, X.; Zhang, L.-P.; Zhang, L.; Yan, P.; Ahammed, G.J.; Han, W.-Y. Methyl salicylate enhances flavonoid biosynthesis in tea leaves by stimulating the phenylpropanoid pathway. Molecules 2019, 24, 362. [Google Scholar] [CrossRef] [Green Version]
  74. Creelman, R.A.; Tierney, M.L.; Mullet, J.E. Jasmonic acid/methyl jasmonate accumulate in wounded soybean hypocotyls and modulate wound gene expression. Proc. Natl. Acad. Sci. USA 1992, 89, 4938–4941. [Google Scholar] [CrossRef] [Green Version]
  75. Sharma, A.; Shahzad, B.; Rehman, A.; Bhardwaj, R.; Landi, M.; Zheng, B. Response of phenylpropanoid pathway and the role of polyphenols in plants under abiotic stress. Molecules 2019, 24, 2452. [Google Scholar] [CrossRef] [Green Version]
  76. Rady, M.O.; Semida, W.M.; Abd El-Mageed, T.A.; Hemida, K.A.; Rady, M.M. Up-regulation of antioxidative defense systems by glycine betaine foliar application in onion plants confer tolerance to salinity stress. Sci. Hortic. 2018, 240, 614–622. [Google Scholar] [CrossRef]
  77. Janská, A.; Maršík, P.; Zelenková, S.; Ovesná, J. Cold stress and acclimation–what is important for metabolic adjustment? Plant Biol. 2010, 12, 395–405. [Google Scholar] [CrossRef]
  78. Kaplan, F.; Kopka, J.; Sung, D.Y.; Zhao, W.; Popp, M.; Porat, R.; Guy, C.L. Transcript and metabolite profiling during cold acclimation of Arabidopsis reveals an intricate relationship of cold-regulated gene expression with modifications in metabolite content. Plant J. 2007, 50, 967–981. [Google Scholar] [CrossRef]
  79. Johnson, H.E.; Broadhurst, D.; Goodacre, R.; Smith, A.R. Metabolic fingerprinting of salt-stressed tomatoes. Phytochemistry 2003, 62, 919–928. [Google Scholar] [CrossRef]
  80. Brosché, M.; Vinocur, B.; Alatalo, E.R.; Lamminmäki, A.; Teichmann, T.; Ottow, E.A.; Djilianov, D.; Afif, D.; Bogeat-Triboulot, M.-B.; Altman, A. Gene expression and metabolite profiling of Populus euphratica growing in the Negev desert. Genome Biol. 2005, 6, 1–17. [Google Scholar] [CrossRef] [Green Version]
  81. Bossolini, E.; Wicker, T.; Knobel, P.A.; Keller, B. Comparison of orthologous loci from small grass genomes Brachypodium and rice: Implications for wheat genomics and grass genome annotation. Plant J. 2007, 49, 704–717. [Google Scholar] [CrossRef] [PubMed]
  82. Gong, Q.; Li, P.; Ma, S.; Indu Rupassara, S.; Bohnert, H.J. Salinity stress adaptation competence in the extremophile Thellungiella halophila in comparison with its relative Arabidopsis thaliana. Plant J. 2005, 44, 826–839. [Google Scholar] [CrossRef] [PubMed]
  83. Kim, J.K.; Bamba, T.; Harada, K.; Fukusaki, E.; Kobayashi, A. Time-course metabolic profiling in Arabidopsis thaliana cell cultures after salt stress treatment. J. Exp. Bot. 2007, 58, 415–424. [Google Scholar] [CrossRef] [PubMed]
  84. Rizhsky, L.; Liang, H.; Shuman, J.; Shulaev, V.; Davletova, S.; Mittler, R. When defense pathways collide. The response of Arabidopsis to a combination of drought and heat stress. Plant Physiol. 2004, 134, 1683–1696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Nikiforova, V.J.; Kopka, J.; Tolstikov, V.; Fiehn, O.; Hopkins, L.; Hawkesford, M.J.; Hesse, H.; Hoefgen, R. Systems rebalancing of metabolism in response to sulfur deprivation, as revealed by metabolome analysis of Arabidopsis plants. Plant Physiol. 2005, 138, 304–318. [Google Scholar] [CrossRef] [Green Version]
  86. Hernández, G.; Ramírez, M.; Valdés-López, O.; Tesfaye, M.; Graham, M.A.; Czechowski, T.; Schlereth, A.; Wandrey, M.; Erban, A.; Cheung, F. Phosphorus stress in common bean: Root transcript and metabolic responses. Plant Physiol. 2007, 144, 752–767. [Google Scholar] [CrossRef] [Green Version]
  87. Morinaka, Y.; Sakamoto, T.; Inukai, Y.; Agetsuma, M.; Kitano, H.; Ashikari, M.; Matsuoka, M. Morphological alteration caused by brassinosteroid insensitivity increases the biomass and grain production of rice. Plant Physiol. 2006, 141, 924–931. [Google Scholar] [CrossRef] [Green Version]
  88. Le Lay, P.; Isaure, M.-P.; Sarry, J.-E.; Kuhn, L.; Fayard, B.; Le Bail, J.-L.; Bastien, O.; Garin, J.; Roby, C.; Bourguignon, J. Metabolomic, proteomic and biophysical analyses of Arabidopsis thaliana cells exposed to a caesium stress. Influence of potassium supply. Biochimie 2006, 88, 1533–1547. [Google Scholar] [CrossRef]
  89. Xu, N.; Fan, X.; Yan, X.; Li, X.; Niu, R.; Tseng, C. Antibacterial bromophenols from the marine red alga Rhodomela confervoides. Phytochemistry 2003, 62, 1221–1224. [Google Scholar] [CrossRef] [PubMed]
  90. Lu, Y.; Lam, H.; Pi, E.; Zhan, Q.; Tsai, S.; Wang, C.; Kwan, Y.; Ngai, S. Comparative metabolomics in Glycine max and Glycine soja under salt stress to reveal the phenotypes of their offspring. J. Agric. Food Chem. 2013, 61, 8711–8721. [Google Scholar] [CrossRef] [PubMed]
  91. Ahsan, N.; Nakamura, T.; Komatsu, S. Differential responses of microsomal proteins and metabolites in two contrasting cadmium (Cd)-accumulating soybean cultivars under Cd stress. Amino Acids 2012, 42, 317–327. [Google Scholar] [CrossRef] [PubMed]
  92. Heinemann, B.; Hildebrandt, T.M. The role of amino acid metabolism in signaling and metabolic adaptation to stress-induced energy deficiency in plants. J. Exp. Bot. 2021, 72, 4634–4645. [Google Scholar] [CrossRef]
  93. Khan, A.; Ali, S.; Murad, W.; Hayat, K.; Siraj, S.; Jawad, M.; Khan, R.A.; Uddin, J.; Al-Harrasi, A.; Khan, A. Phytochemical and pharmacological uses of medicinal plants to treat cancer: A case study from Khyber Pakhtunkhwa, North Pakistan. J. Ethnopharmacol. 2021, 281, 114437. [Google Scholar] [CrossRef] [PubMed]
  94. Hildebrandt, T.M. Synthesis versus degradation: Directions of amino acid metabolism during Arabidopsis abiotic stress response. Plant Mol. Biol. 2018, 98, 121–135. [Google Scholar] [CrossRef] [PubMed]
  95. Khan, N.; Ali, S.; Zandi, P.; Mehmood, A.; Ullah, S.; Ikram, M.; Ismail, M.A.S.; Babar, M. Role of sugars, amino acids and organic acids in improving plant abiotic stress tolerance. Pak. J. Bot 2020, 52, 355–363. [Google Scholar] [CrossRef]
  96. Rodziewicz, P.; Swarcewicz, B.; Chmielewska, K.; Wojakowska, A.; Stobiecki, M. Influence of abiotic stresses on plant proteome and metabolome changes. Acta Physiol. Plant. 2014, 36, 1–19. [Google Scholar] [CrossRef] [Green Version]
  97. Verbruggen, N.; Hermans, C. Proline accumulation in plants: A review. Amino Acids 2008, 35, 753–759. [Google Scholar] [CrossRef]
  98. Meena, M.; Divyanshu, K.; Kumar, S.; Swapnil, P.; Zehra, A.; Shukla, V.; Yadav, M.; Upadhyay, R. Regulation of L-proline biosynthesis, signal transduction, transport, accumulation and its vital role in plants during variable environmental conditions. Heliyon 2019, 5, e02952. [Google Scholar] [CrossRef] [Green Version]
  99. Yang, M.M.; Wang, J.; Dong, L.; Kong, D.J.; Teng, Y.; Liu, P.; Fan, J.J.; Yu, X.H. Lack of association of C3 gene with uveitis: Additional insights into the genetic profile of uveitis regarding complement pathway genes. Sci. Rep. 2017, 7, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Hanif, S.; Saleem, M.F.; Sarwar, M.; Irshad, M.; Shakoor, A.; Wahid, M.A.; Khan, H.Z. Biochemically triggered heat and drought stress tolerance in rice by proline application. J. Plant Growth Regul. 2021, 40, 305–312. [Google Scholar] [CrossRef]
  101. Rashedy, A.A.; Abd-ElNafea, M.H.; Khedr, E.H. Co-application of proline or calcium and humic acid enhances productivity of salt stressed pomegranate by improving nutritional status and osmoregulation mechanisms. Sci. Rep. 2022, 12, 1–10. [Google Scholar] [CrossRef] [PubMed]
  102. Abeed, A.H.; Eissa, M.A.; Abdel-Wahab, D.A. Effect of exogenously applied jasmonic acid and kinetin on drought tolerance of wheat cultivars based on morpho-physiological evaluation. J. Soil Sci. Plant Nutr. 2021, 21, 131–144. [Google Scholar] [CrossRef]
  103. Vasquez-Robinet, C.; Mane, S.P.; Ulanov, A.V.; Watkinson, J.I.; Stromberg, V.K.; De Koeyer, D.; Schafleitner, R.; Willmot, D.B.; Bonierbale, M.; Bohnert, H.J. Physiological and molecular adaptations to drought in Andean potato genotypes. J. Exp. Bot. 2008, 59, 2109–2123. [Google Scholar] [CrossRef] [Green Version]
  104. Li, Q.; Zhao, Y.; Ding, W.; Han, B.; Geng, S.; Ning, D.; Ma, T.; Yu, X. Gamma-aminobutyric acid facilitates the simultaneous production of biomass, astaxanthin and lipids in Haematococcus pluvialis under salinity and high-light stress conditions. Bioresour. Technol. 2021, 320, 124418. [Google Scholar] [CrossRef]
  105. Lv, W.-T.; Lin, B.; Zhang, M.; Hua, X.-J. Proline accumulation is inhibitory to Arabidopsis seedlings during heat stress. Plant Physiol. 2011, 156, 1921–1933. [Google Scholar] [CrossRef] [Green Version]
  106. Charlton, A.J.; Donarski, J.A.; Harrison, M.; Jones, S.A.; Godward, J.; Oehlschlager, S.; Arques, J.L.; Ambrose, M.; Chinoy, C.; Mullineaux, P.M. Responses of the pea (Pisum sativum L.) leaf metabolome to drought stress assessed by nuclear magnetic resonance spectroscopy. Metabolomics 2008, 4, 312–327. [Google Scholar] [CrossRef]
  107. Kalaji, H.M.; Jajoo, A.; Oukarroum, A.; Brestic, M.; Zivcak, M.; Samborska, I.A.; Cetner, M.D.; Łukasik, I.; Goltsev, V.; Ladle, R.J. Chlorophyll a fluorescence as a tool to monitor physiological status of plants under abiotic stress conditions. Acta Physiol. Plant. 2016, 38, 102. [Google Scholar] [CrossRef] [Green Version]
  108. Mao, Y.-B.; Liu, Y.-Q.; Chen, D.-Y.; Chen, F.-Y.; Fang, X.; Hong, G.-J.; Wang, L.-J.; Wang, J.-W.; Chen, X.-Y. Jasmonate response decay and defense metabolite accumulation contributes to age-regulated dynamics of plant insect resistance. Nat. Commun. 2017, 8, 13925 . [Google Scholar] [CrossRef] [Green Version]
  109. Liu, J.-H.; Wang, W.; Wu, H.; Gong, X.; Moriguchi, T. Polyamines function in stress tolerance: From synthesis to regulation. Front. Plant Sci. 2015, 6, 827. [Google Scholar] [CrossRef] [Green Version]
  110. Fischer, W.; Calderón, M.; Haag, R. Hyperbranched polyamines for transfection. Nucleic Acid Transfection 2010, 95–129. [Google Scholar]
  111. Chen, D.; Shao, Q.; Yin, L.; Younis, A.; Zheng, B. Polyamine function in plants: Metabolism, regulation on development, and roles in abiotic stress responses. Front. Plant Sci. 2019, 9, 1945. [Google Scholar] [CrossRef]
  112. Alcázar, R.; Bueno, M.; Tiburcio, A.F. Polyamines: Small amines with large effects on plant abiotic stress tolerance. Cells 2020, 9, 2373. [Google Scholar] [CrossRef] [PubMed]
  113. Capell, T.; Bassie, L.; Christou, P. Modulation of the polyamine biosynthetic pathway in transgenic rice confers tolerance to drought stress. Proc. Natl. Acad. Sci. USA 2004, 101, 9909–9914. [Google Scholar] [CrossRef] [Green Version]
  114. Zhao, J.; Wang, X.; Pan, X.; Jiang, Q.; Xi, Z. Exogenous putrescine alleviates drought stress by altering reactive oxygen species scavenging and biosynthesis of polyamines in the seedlings of Cabernet Sauvignon. Front. Plant Sci. 2021, 12. [Google Scholar] [CrossRef] [PubMed]
  115. Vickers, N.J. Animal communication: When i’m calling you, will you answer too? Curr. Biol. 2017, 27, R713–R715. [Google Scholar] [CrossRef]
  116. Gill, S.S.; Tuteja, N. Polyamines and abiotic stress tolerance in plants. Plant Signal. Behav. 2010, 5, 26–33. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  117. Saxena, R.; Adhikari, D.; Goyal, H. Biomass-based energy fuel through biochemical routes: A review. Renew. Sustain. Energy Rev. 2009, 13, 167–178. [Google Scholar] [CrossRef]
  118. Ruppert, A.M.; Weinberg, K.; Palkovits, R. Hydrogenolysis goes bio: From carbohydrates and sugar alcohols to platform chemicals. Angew. Chem. Int. Ed. 2012, 51, 2564–2601. [Google Scholar] [CrossRef]
  119. Cao, X.; Zhu, C.; Zhong, C.; Hussain, S.; Zhu, L.; Wu, L.; Jin, Q. Mixed-nitrogen nutrition-mediated enhancement of drought tolerance of rice seedlings associated with photosynthesis, hormone balance and carbohydrate partitioning. Plant Growth Regul. 2018, 84, 451–465. [Google Scholar] [CrossRef]
  120. Rosa, M.; Prado, C.; Podazza, G.; Interdonato, R.; González, J.A.; Hilal, M.; Prado, F.E. Soluble sugars: Metabolism, sensing and abiotic stress: A complex network in the life of plants. Plant Signal. Behav. 2009, 4, 388–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Sicher, R.C.; Timlin, D.; Bailey, B. Responses of growth and primary metabolism of water-stressed barley roots to rehydration. J. Plant Physiol. 2012, 169, 686–695. [Google Scholar] [CrossRef]
  122. Sperdouli, I.; Moustakas, M. Interaction of proline, sugars, and anthocyanins during photosynthetic acclimation of Arabidopsis thaliana to drought stress. J. Plant Physiol. 2012, 169, 577–585. [Google Scholar] [CrossRef]
  123. Basu, P.; Ali, M.; Chaturvedi, S. Osmotic adjustment increases water uptake, remobilization of assimilates and maintains photosynthesis in chickpea under drought. Experiment 2007, 45, 261–267. [Google Scholar]
  124. Kaplan, F.; Guy, C.L. β-Amylase induction and the protective role of maltose during temperature shock. Plant Physiol. 2004, 135, 1674–1684. [Google Scholar] [CrossRef] [Green Version]
  125. Keunen, E.; Peshev, D.; Vangronsveld, J.; Van Den Ende, W.; Cuypers, A. Plant sugars are crucial players in the oxidative challenge during abiotic stress: Extending the traditional concept. Plant Cell Environ. 2013, 36, 1242–1255. [Google Scholar] [CrossRef]
  126. Saddhe, A.A.; Manuka, R.; Penna, S. Plant sugars: Homeostasis and transport under abiotic stress in plants. Physiol. Plant. 2021, 171, 739–755. [Google Scholar] [CrossRef]
  127. Krasavina, M.S.; Burmistrova, N.A.; Raldugina, G.N. The role of carbohydrates in plant resistance to abiotic stresses. In Emerging Technologies and Management of Crop Stress Tolerance; Elsevier: Amsterdam, The Netherlands, 2014; pp. 229–270. [Google Scholar]
  128. Raza, S.H.; Athar, H.R.; Ashraf, M.; Hameed, A. Glycinebetaine-induced modulation of antioxidant enzymes activities and ion accumulation in two wheat cultivars differing in salt tolerance. Environ. Exp. Bot. 2007, 60, 368–376. [Google Scholar] [CrossRef]
  129. Chen, T.H.; Murata, N. Glycinebetaine: An effective protectant against abiotic stress in plants. Trends Plant Sci. 2008, 13, 499–505. [Google Scholar] [CrossRef] [PubMed]
  130. Giri, J. Glycinebetaine and abiotic stress tolerance in plants. Plant Signal. Behav. 2011, 6, 1746–1751. [Google Scholar] [CrossRef] [PubMed]
  131. Sakamotto, A.; Muratta, N. The role of glycine betaine in protection of plants against stress: Clue from transgenic plants. Plant Cell Environ. 2002, 25, 163–171. [Google Scholar] [CrossRef] [PubMed]
  132. Quan, R.; Shang, M.; Zhang, H.; Zhao, Y.; Zhang, J. Improved chilling tolerance by transformation with betA gene for the enhancement of glycinebetaine synthesis in maize. Plant Sci. 2004, 166, 141–149. [Google Scholar] [CrossRef]
  133. Prasad, K.; Sharmila, P.; Kumar, P.; Saradhi, P.P. Transformation of Brassica juncea (L.) Czern with bacterial codA gene enhances its tolerance to salt stress. Mol. Breed. 2000, 6, 489–499. [Google Scholar] [CrossRef]
  134. Holmström, K.O.; Somersalo, S.; Mandal, A.; Palva, T.E.; Welin, B. Improved tolerance to salinity and low temperature in transgenic tobacco producing glycine betaine. J. Exp. Bot. 2000, 51, 177–185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Niesche, R.; Haase, M. Emotions and ethics: A Foucauldian framework for becoming an ethical educator. Educ. Philos. Theory 2012, 44, 276–288. [Google Scholar] [CrossRef]
  136. Singh, D.; Singh, C.K.; Singh, D.; Sarkar, S.K.; Prasad, S.K.; Sharma, N.L.; Singh, I. Glycine betaine modulates chromium (VI)-induced morpho-physiological and biochemical responses to mitigate chromium toxicity in chickpea (Cicer arietinum L.) cultivars. Sci. Rep. 2022, 12, 8005. [Google Scholar]
  137. Khedr, R.A.; Sorour, S.G.R.; Aboukhadrah, S.H.; El Shafey, N.M.; Abd Elsalam, H.E.; El-Sharnouby, M.E.; El-Tahan, A.M. Alleviation of salinity stress effects on agro-physiological traits of wheat by auxin, glycine betaine, and soil additives. Saudi J. Biol. Sci. 2022, 29, 534–540. [Google Scholar] [CrossRef]
  138. Emamverdian, A.; Ding, Y.; Mokhberdoran, F.; Xie, Y. Heavy metal stress and some mechanisms of plant defense response. Sci. World J. 2015, 2015, 756120 . [Google Scholar] [CrossRef] [Green Version]
  139. Fahy, E.; Subramaniam, S.; Brown, H.A.; Glass, C.K.; Merrill, A.H.; Murphy, R.C.; Raetz, C.R.; Russell, D.W.; Seyama, Y.; Shaw, W. A comprehensive classification system for lipids1. J. Lipid Res. 2005, 46, 839–861. [Google Scholar] [CrossRef] [Green Version]
  140. Okazaki, Y.; Saito, K. Roles of lipids as signaling molecules and mitigators during stress response in plants. Plant J. 2014, 79, 584–596. [Google Scholar] [CrossRef]
  141. Munnik, T.; Zarza, X. Analyzing plant signaling phospholipids through 32 P i-labeling and TLC. In Plant Lipid Signaling Protocols; Springer: Berlin/Heidelberg, Germany, 2013; pp. 3–15. [Google Scholar]
  142. Welti, R.; Li, W.; Li, M.; Sang, Y.; Biesiada, H.; Zhou, H.-E.; Rajashekar, C.; Williams, T.D.; Wang, X. Profiling membrane lipids in plant stress responses: Role of phospholipase Dα in freezing-induced lipid changes in Arabidopsis. J. Biol. Chem. 2002, 277, 31994–32002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Viehweger, K.; Dordschbal, B.; Roos, W. Elicitor-activated phospholipase A2 generates lysophosphatidylcholines that mobilize the vacuolar H+ pool for pH signaling via the activation of Na+-dependent proton fluxes. Plant Cell 2002, 14, 1509–1525. [Google Scholar] [CrossRef] [Green Version]
  144. Drissner, D.; Kunze, G.; Callewaert, N.; Gehrig, P.; Tamasloukht, M.B.; Boller, T.; Felix, G.; Amrhein, N.; Bucher, M. Lyso-phosphatidylcholine is a signal in the arbuscular mycorrhizal symbiosis. Science 2007, 318, 265–268. [Google Scholar] [CrossRef] [PubMed]
  145. Mandal, M.K.; Chandra-Shekara, A.; Jeong, R.-D.; Yu, K.; Zhu, S.; Chanda, B.; Navarre, D.; Kachroo, A.; Kachroo, P. Oleic acid–dependent modulation of NITRIC OXIDE ASSOCIATED1 protein levels regulates nitric oxide–mediated defense signaling in Arabidopsis. Plant Cell 2012, 24, 1654–1674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  146. Upchurch, R.G. Fatty acid unsaturation, mobilization, and regulation in the response of plants to stress. Biotechnol. Lett. 2008, 30, 967–977. [Google Scholar] [CrossRef]
  147. Markham, J.E.; Lynch, D.V.; Napier, J.A.; Dunn, T.M.; Cahoon, E.B. Plant sphingolipids: Function follows form. Curr. Opin. Plant Biol. 2013, 16, 350–357. [Google Scholar] [CrossRef]
  148. Ballaré, C.L. Jasmonate-induced defenses: A tale of intelligence, collaborators and rascals. Trends Plant Sci. 2011, 16, 249–257. [Google Scholar] [CrossRef]
  149. Kang, L.; Wang, Y.S.; Uppalapati, S.R.; Wang, K.; Tang, Y.; Vadapalli, V.; Venables, B.J.; Chapman, K.D.; Blancaflor, E.B.; Mysore, K.S. Overexpression of a fatty acid amide hydrolase compromises innate immunity in Arabidopsis. Plant J. 2008, 56, 336–349. [Google Scholar] [CrossRef]
  150. Sun, Y.; Li, Y.; Sun, X.; Wu, Q.; Yang, C.; Wang, L. Overexpression of a Phosphatidylinositol-Specific Phospholipase C Gene from Populus simonii× P. nigra Improves Salt Tolerance in Transgenic Tobacco. J. Plant Biol. 2022, 65, 365–376. [Google Scholar] [CrossRef]
  151. Pinosa, F.; Buhot, N.; Kwaaitaal, M.; Fahlberg, P.; Thordal-Christensen, H.; Ellerström, M.; Andersson, M.X. Arabidopsis phospholipase Dδ is involved in basal defense and nonhost resistance to powdery mildew fungi. Plant Physiol. 2013, 163, 896–906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Arisz, S.A.; van Wijk, R.; Roels, W.; Zhu, J.-K.; Haring, M.A.; Munnik, T. Rapid phosphatidic acid accumulation in response to low temperature stress in Arabidopsis is generated through diacylglycerol kinase. Front. Plant Sci. 2013, 4, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. McLoughlin, F.; Arisz, S.A.; Dekker, H.L.; Kramer, G.; De Koster, C.G.; Haring, M.A.; Munnik, T.; Testerink, C. Identification of novel candidate phosphatidic acid-binding proteins involved in the salt-stress response of Arabidopsis thaliana roots. Biochem. J. 2013, 450, 573–581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Uraji, M.; Katagiri, T.; Okuma, E.; Ye, W.; Hossain, M.A.; Masuda, C.; Miura, A.; Nakamura, Y.; Mori, I.C.; Shinozaki, K. Cooperative function of PLDδ and PLDα1 in abscisic acid-induced stomatal closure in Arabidopsis. Plant Physiol. 2012, 159, 450–460. [Google Scholar] [CrossRef] [Green Version]
  155. Wink, M. Introduction: Biochemistry, physiology and ecological functions of secondary metabolites. In Annual Plant Reviews Volume 40: Biochemistry of Plant Secondary Metabolism; Wiley: Hoboken, NJ, USA, 2010; pp. 1–19. [Google Scholar]
  156. Böttger, A.; Vothknecht, U.; Bolle, C.; Wolf, A. Lessons on Caffeine, Cannabis & Co.; Springer: Cham, Switzerland, 2018; pp. 171–178. [Google Scholar]
  157. Debona, D.; Rodrigues, F.A.; Datnoff, L.E. Silicon's role in abiotic and biotic plant stresses. Annu. Rev. Phytopathol. 2017, 55, 85–107. [Google Scholar] [CrossRef] [Green Version]
  158. Divekar, P.A.; Narayana, S.; Divekar, B.A.; Kumar, R.; Gadratagi, B.G.; Ray, A.; Singh, A.K.; Rani, V.; Singh, V.; Singh, A.K. Plant secondary metabolites as defense tools against herbivores for sustainable crop protection. Int. J. Mol. Sci. 2022, 23, 2690. [Google Scholar] [CrossRef]
  159. Ochoa-Villarreal, M.; Howat, S.; Hong, S.; Jang, M.O.; Jin, Y.-W.; Lee, E.-K.; Loake, G.J. Plant cell culture strategies for the production of natural products. BMB Rep. 2016, 49, 149. [Google Scholar] [CrossRef] [PubMed]
  160. Edreva, A.; Velikova, V.; Tsonev, T. Phenylamides in plants. Russ. J. Plant Physiol. 2007, 54, 287–301. [Google Scholar] [CrossRef]
  161. Pagare, S.; Bhatia, M.; Tripathi, N.; Pagare, S.; Bansal, Y. Secondary metabolites of plants and their role: Overview. Curr. Trends Biotechnol. Pharm. 2015, 9, 293–304. [Google Scholar]
  162. Elateeq, A.A.; Ahmed, M.; Abdelkawy, A.M.; Toaima, N.M.; Bosila, H.A.; Zarad, M.M.; Ebrahim, H.S.; Jiao, J.; Hongyi, P.; Ullah, S. Establishment of Gypsophila paniculata root culture for biomass, saponin, and flavonoid production. Not. Bot. Horti Agrobot. Cluj-Napoca 2022, 50, 12886. [Google Scholar] [CrossRef]
  163. Bosila, H.; Hamza, M.A.; El-Ateeq, A. Enhancement of callus growth and hyoscyamine alkaloid production in Hyoscyamus muticus by nanotechnology, biotic elicitor and precursor. Int. J. ChemTech. Res. 2016, 9, 135–142. [Google Scholar]
  164. Ali, M.S.; Baek, K.-H. Jasmonic acid signaling pathway in response to abiotic stresses in plants. Int. J. Mol. Sci. 2020, 21, 621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Elshahawy, O.A.; Zeawail, M.E.-F.; Hamza, M.A.; Elateeq, A.A.; Omar, M.A. Improving the Production of Total Phenolics and Flavonoids and the Antioxidant Capacity of Echinacea purpurea Callus through Biotic Elicitation. Egypt. J. Chem. 2022, 65, 137–149. [Google Scholar] [CrossRef]
  166. Mahdi, J.; Mahdi, A.; Mahdi, A.; Bowen, I. The historical analysis of aspirin discovery, its relation to the willow tree and antiproliferative and anticancer potential. Cell Prolif. 2006, 39, 147–155. [Google Scholar] [CrossRef] [PubMed]
  167. Yazici, L. Influence of different sowing times on yield and biochemical characteristics of different opium poppy (Papaver somniferum L.) genotypes. J. King Saud Univ.-Sci. 2022, 34, 102337. [Google Scholar] [CrossRef]
  168. Grace, S.C. Phenolics as antioxidants. Antioxid. React. Oxyg. Species Plants 2005, 141, 168. [Google Scholar]
  169. Waterman, P.G. Roles for secondary metabolites in plants. In Ciba Foundation Symposium 171-Secondary Metabolites: Their Function and Evolution: Secondary Metabolites: Their Function and Evolution: Ciba Foundation Symposium 171; Wiley Online Library: Hoboken, NJ, USA, 2007. [Google Scholar]
  170. Kabera, J.N.; Semana, E.; Mussa, A.R.; He, X. Plant secondary metabolites: Biosynthesis, classification, function and pharmacological properties. J. Pharm. Pharm. 2014, 2, 377–392. [Google Scholar]
  171. Saltveit, M.E. Synthesis and metabolism of phenolic compounds. Fruit Veget. Phytochem. Chem. Hum. Health 2017, 2, 115. [Google Scholar]
  172. Dong, N.Q.; Lin, H.X. Contribution of phenylpropanoid metabolism to plant development and plant–environment interactions. J. Integr. Plant Biol. 2021, 63, 180–209. [Google Scholar] [CrossRef]
  173. Valifard, M.; Mohsenzadeh, S.; Niazi, A.; Moghadam, A. Phenylalanine ammonia lyase isolation and functional analysis of phenylpropanoid pathway under salinity stress in'Salvia'species. Aust. J. Crop Sci. 2015, 9, 656–665. [Google Scholar]
  174. Amarowicz, R.; Weidner, S. Phenolic compounds and properties of antioxidants in grapevine roots (Vitis vinifera L.) under drought stress followed by recovery. Acta Soc. Bot. Pol. 2009, 78, 389–405. [Google Scholar]
  175. Nada, R.S.; Ashmawi, A.E.; Mady, E.; Randhir, T.O.; Elateeq, A.A. Effect of Organic Manure and Plant Growth Promoting Microbes on Yield, Quality and Essential Oil Constituents of Fennel Bulb (Foeniculum vulgare Mill.). J. Ecol. Eng. 2022, 23, 149–164. [Google Scholar] [CrossRef]
  176. Cheynier, V.; Comte, G.; Davies, K.M.; Lattanzio, V.; Martens, S. Plant phenolics: Recent advances on their biosynthesis, genetics, and ecophysiology. Plant Physiol. Biochem. 2013, 72, 1–20. [Google Scholar] [CrossRef] [PubMed]
  177. Kuluev, B.; Mikhaylova, E.; Berezhneva, Z.; Nikonorov, Y.; Postrigan, B.; Kudoyarova, G.; Chemeris, A. Expression profiles and hormonal regulation of tobacco NtEXGT gene and its involvement in abiotic stress response. Plant Physiol. Biochem. 2017, 111, 203–215. [Google Scholar] [CrossRef]
  178. Signorelli, S.; Coitiño, E.L.; Borsani, O.; Monza, J. Molecular mechanisms for the reaction between OH radicals and proline: Insights on the role as reactive oxygen species scavenger in plant stress. J. Phys. Chem. B 2014, 118, 37–47. [Google Scholar] [CrossRef] [PubMed]
  179. Dehghan, S.; Sadeghi, M.; Pöppel, A.; Fischer, R.; Lakes-Harlan, R.; Kavousi, H.R.; Vilcinskas, A.; Rahnamaeian, M. Differential inductions of phenylalanine ammonia-lyase and chalcone synthase during wounding, salicylic acid treatment, and salinity stress in safflower, Carthamus tinctorius. Biosci. Rep. 2014, 34, 273–282. [Google Scholar] [CrossRef]
  180. Hernandez-Aguilar, C.; Dominguez-Pacheco, A.; Tenango, M.P.; Valderrama-Bravo, C.; Hernández, M.S.; Cruz-Orea, A.; Ordonez-Miranda, J. Characterization of bean seeds, germination, and phenolic compounds of seedlings by UV-C radiation. J. Plant Growth Regul. 2021, 40, 642–655. [Google Scholar] [CrossRef]
  181. Mencin, M.; Abramovič, H.; Jamnik, P.; Petkovšek, M.M.; Veberič, R.; Terpinc, P. Abiotic stress combinations improve the phenolics profiles and activities of extractable and bound antioxidants from germinated spelt (Triticum spelta L.) seeds. Food Chem. 2021, 344, 128704. [Google Scholar] [CrossRef]
  182. Huihui, Z.; Xin, L.; Zisong, X.; Yue, W.; Zhiyuan, T.; Meijun, A.; Yuehui, Z.; Wenxu, Z.; Nan, X.; Guangyu, S. Toxic effects of heavy metals Pb and Cd on mulberry (Morus alba L.) seedling leaves: Photosynthetic function and reactive oxygen species (ROS) metabolism responses. Ecotoxicol. Environ. Saf. 2020, 195, 110469. [Google Scholar] [CrossRef]
  183. Kohli, S.K.; Handa, N.; Sharma, A.; Gautam, V.; Arora, S.; Bhardwaj, R.; Wijaya, L.; Alyemeni, M.N.; Ahmad, P. Interaction of 24-epibrassinolide and salicylic acid regulates pigment contents, antioxidative defense responses, and gene expression in Brassica juncea L. seedlings under Pb stress. Environ. Sci. Pollut. Res. 2018, 25, 15159–15173. [Google Scholar] [CrossRef]
  184. Handa, N.; Kohli, S.K.; Sharma, A.; Thukral, A.K.; Bhardwaj, R.; Abd_Allah, E.F.; Alqarawi, A.A.; Ahmad, P. Selenium modulates dynamics of antioxidative defence expression, photosynthetic attributes and secondary metabolites to mitigate chromium toxicity in Brassica juncea L. plants. Environ. Exp. Bot. 2019, 161, 180–192. [Google Scholar] [CrossRef]
  185. Alvarez, S.; Marsh, E.L.; Schroeder, S.G.; Schachtman, D.P. Metabolomic and proteomic changes in the xylem sap of maize under drought. Plant Cell Environ. 2008, 31, 325–340. [Google Scholar] [CrossRef] [PubMed]
  186. Fan, L.; Linker, R.; Gepstein, S.; Tanimoto, E.; Yamamoto, R.; Neumann, P.M. Progressive inhibition by water deficit of cell wall extensibility and growth along the elongation zone of maize roots is related to increased lignin metabolism and progressive stelar accumulation of wall phenolics. Plant Physiol. 2006, 140, 603–612. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Li, B.; Fan, R.; Sun, G.; Sun, T.; Fan, Y.; Bai, S.; Guo, S.; Huang, S.; Liu, J.; Zhang, H. Flavonoids improve drought tolerance of maize seedlings by regulating the homeostasis of reactive oxygen species. Plant Soil 2021, 461, 389–405. [Google Scholar] [CrossRef]
  188. Nakabayashi, R.; Yonekura-Sakakibara, K.; Urano, K.; Suzuki, M.; Yamada, Y.; Nishizawa, T.; Matsuda, F.; Kojima, M.; Sakakibara, H.; Shinozaki, K. Enhancement of oxidative and drought tolerance in Arabidopsis by overaccumulation of antioxidant flavonoids. Plant J. 2014, 77, 367–379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Nichols, S.N.; Hofmann, R.W.; Williams, W.M. Physiological drought resistance and accumulation of leaf phenolics in white clover interspecific hybrids. Environ. Exp. Bot. 2015, 119, 40–47. [Google Scholar] [CrossRef]
  190. André, C.M.; Schafleitner, R.; Legay, S.; Lefèvre, I.; Aliaga, C.A.A.; Nomberto, G.; Hoffmann, L.; Hausman, J.-F.; Larondelle, Y.; Evers, D. Gene expression changes related to the production of phenolic compounds in potato tubers grown under drought stress. Phytochemistry 2009, 70, 1107–1116. [Google Scholar] [CrossRef]
  191. Parida, A.K.; Dagaonkar, V.S.; Phalak, M.S.; Umalkar, G.; Aurangabadkar, L.P. Alterations in photosynthetic pigments, protein and osmotic components in cotton genotypes subjected to short-term drought stress followed by recovery. Plant Biotechnol. Rep. 2007, 1, 37–48. [Google Scholar] [CrossRef]
  192. Naikoo, M.I.; Dar, M.I.; Raghib, F.; Jaleel, H.; Ahmad, B.; Raina, A.; Khan, F.A.; Naushin, F. Role and regulation of plants phenolics in abiotic stress tolerance: An overview. Plant Signal. Mol. 2019, 2, 157–168. [Google Scholar]
  193. de Abreu, I.N.; Mazzafera, P. Effect of water and temperature stress on the content of active constituents of Hypericum brasiliense Choisy. Plant Physiol. Biochem. 2005, 43, 241–248. [Google Scholar] [CrossRef]
  194. Djoukeng, J.D.; Arbona, V.; Argamasilla, R.; Gomez-Cadenas, A. Flavonoid profiling in leaves of citrus genotypes under different environmental situations. J. Agric. Food Chem. 2008, 56, 11087–11097. [Google Scholar] [CrossRef]
  195. Ancillotti, C.; Bogani, P.; Biricolti, S.; Calistri, E.; Checchini, L.; Ciofi, L.; Gonnelli, C.; Del Bubba, M. Changes in polyphenol and sugar concentrations in wild type and genetically modified Nicotiana langsdorffii Weinmann in response to water and heat stress. Plant Physiol. Biochem. 2015, 97, 52–61. [Google Scholar] [CrossRef]
  196. Xiang, M.; Ding, W.; Wu, C.; Wang, W.; Ye, S.; Cai, C.; Hu, X.; Wang, N.; Bai, W.; Tang, X. Production of purple Ma bamboo (Dendrocalamus latiflorus Munro) with enhanced drought and cold stress tolerance by engineering anthocyanin biosynthesis. Planta 2021, 254, 50 . [Google Scholar] [CrossRef] [PubMed]
  197. Wang, N.; Zhang, Z.; Jiang, S.; Xu, H.; Wang, Y.; Feng, S.; Chen, X. Synergistic effects of light and temperature on anthocyanin biosynthesis in callus cultures of red-fleshed apple (Malus sieversii f. niedzwetzkyana). Plant Cell Tissue Organ Cult. (PCTOC) 2016, 127, 217–227. [Google Scholar] [CrossRef] [Green Version]
  198. Cabane, M.; Afif, D.; Hawkins, S. Lignins and abiotic stresses. In Advances in Botanical Research; Elsevier: Amsterdam, The Netherlands, 2012; pp. 219–262. [Google Scholar]
  199. Bulotta, S.; Oliverio, M.; Russo, D.; Procopio, A. Biological Activity of Oleuropein and Its Derivatives; Springer: Berlin/Heidelberg, Germany, 2013; pp. 3605–3638. [Google Scholar]
  200. Elateeq, A.A.; Sun, Y.; Nxumalo, W.; Gabr, A.M. Biotechnological production of silymarin in Silybum marianum L.: A review. Biocatal. Agric. Biotechnol. 2020, 29, 101775. [Google Scholar] [CrossRef]
  201. Elateeq, A.A.; Saad, Z.; Eissa, M.; Ullah, S. Effect of chitosan and light conditions on the production of callus biomass, total flavonoids and total phenolics in Ginkgo biloba L. Al-Azhar J. Agric. Res. 2021, 46, 28–42. [Google Scholar] [CrossRef]
  202. Mujib, A.; Fatima, S.; Malik, M.Q. Gamma ray–induced tissue responses and improved secondary metabolites accumulation in Catharanthus roseus. Appl. Microbiol. Biotechnol. 2022, 106, 6109–6123. [Google Scholar] [CrossRef] [PubMed]
  203. Andersen, E.J.; Ali, S.; Byamukama, E.; Yen, Y.; Nepal, M.P. Disease resistance mechanisms in plants. Genes 2018, 9, 339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Toffolatti, S.L.; Maddalena, G.; Passera, A.; Casati, P.; Bianco, P.A.; Quaglino, F. Role of terpenes in plant defense to biotic stress. In Biocontrol Agents and Secondary Metabolites; Elsevier: Amsterdam, The Netherlands, 2021; pp. 401–417. [Google Scholar]
  205. Kumar, S.; Shah, S.H.; Vimala, Y.; Jatav, H.S.; Ahmad, P.; Chen, Y.; Siddique, K.H. Abscisic acid: Metabolism, transport, crosstalk with other plant growth regulators, and its role in heavy metal stress mitigation. Front. Plant Sci. 2022, 13, 972856. [Google Scholar] [CrossRef]
  206. Al Musayeib, N.M.; Musarat, A.; Maqsood, F. Terpenes: A Source of Novel Antimicrobials, Applications and Recent Advances. In Eco-Friendly Biobased Products Used in Microbial Diseases; CRC Press: Boca Raton, FL, USA, 2022; pp. 247–270. [Google Scholar]
  207. Gu, C.-Z.; Xia, X.-M.; Lv, J.; Tan, J.-W.; Baerson, S.R.; Pan, Z.-q.; Song, Y.-Y.; Zeng, R.-S. Diterpenoids with herbicidal and antifungal activities from hulls of rice (Oryza sativa). Fitoterapia 2019, 136, 104183. [Google Scholar] [CrossRef]
  208. Yang, C.-Q.; Wu, X.-M.; Ruan, J.-X.; Hu, W.-L.; Mao, Y.-B.; Chen, X.-Y.; Wang, L.-J. Isolation and characterization of terpene synthases in cotton (Gossypium hirsutum). Phytochemistry 2013, 96, 46–56. [Google Scholar] [CrossRef]
  209. Vaughan, M.M.; Christensen, S.; Schmelz, E.A.; Huffaker, A.; Mcauslane, H.J.; Alborn, H.T.; Romero, M.; Allen, L.H.; Teal, P.E. Accumulation of terpenoid phytoalexins in maize roots is associated with drought tolerance. Plant Cell Environ. 2015, 38, 2195–2207. [Google Scholar] [CrossRef] [PubMed]
  210. Massacci, A.; Nabiev, S.; Pietrosanti, L.; Nematov, S.; Chernikova, T.; Thor, K.; Leipner, J. Response of the photosynthetic apparatus of cotton (Gossypium hirsutum) to the onset of drought stress under field conditions studied by gas-exchange analysis and chlorophyll fluorescence imaging. Plant Physiol. Biochem. 2008, 46, 189–195. [Google Scholar] [CrossRef] [PubMed]
  211. Yusuf, M.A.; Kumar, D.; Rajwanshi, R.; Strasser, R.J.; Tsimilli-Michael, M.; Sarin, N.B. Overexpression of γ-tocopherol methyl transferase gene in transgenic Brassica juncea plants alleviates abiotic stress: Physiological and chlorophyll a fluorescence measurements. Biochim. Et Biophys. Acta (BBA)-Bioenerg. 2010, 1797, 1428–1438. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Wu, W.; Zhang, Q.; Zhu, Y.; Lam, H.-M.; Cai, Z.; Guo, D. Comparative metabolic profiling reveals secondary metabolites correlated with soybean salt tolerance. J. Agric. Food Chem. 2008, 56, 11132–11138. [Google Scholar] [CrossRef] [PubMed]
  213. Martinez, V.; Mestre, T.C.; Rubio, F.; Girones-Vilaplana, A.; Moreno, D.A.; Mittler, R.; Rivero, R.M. Accumulation of flavonols over hydroxycinnamic acids favors oxidative damage protection under abiotic stress. Front. Plant Sci. 2016, 7, 838. [Google Scholar] [CrossRef] [Green Version]
  214. Davodnia, B.; Ahmahdi, J.; Fabriki Ourang, S. Evaluation of drought and salinity stresses on morphological and biochemical characteristics in four species of Papaver. Eco-Phytochem. J. Med. Plants 2017, 5, 24–36. [Google Scholar]
  215. Mewis, I.; Khan, M.A.; Glawischnig, E.; Schreiner, M.; Ulrichs, C. Water stress and aphid feeding differentially influence metabolite composition in Arabidopsis thaliana (L.). PLoS ONE 2012, 7, e48661. [Google Scholar] [CrossRef] [Green Version]
  216. Taha, R.S.; Seleiman, M.F.; Alhammad, B.A.; Alkahtani, J.; Alwahibi, M.S.; Mahdi, A.H. Activated Yeast extract enhances growth, anatomical structure, and productivity of Lupinus termis L. plants under actual salinity conditions. Agronomy 2020, 11, 74. [Google Scholar] [CrossRef]
  217. Miyagawa, H. Studies on nitrogen-containing secondary metabolites playing a defensive role in plants. J. Pestic. Sci. 2009, 34, 110–112. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Respective illustration of the processes involved in plant metabolomics analysis of GC–MS, LC-MS, CE-MS, and NMR-based chromatography.
Figure 1. Respective illustration of the processes involved in plant metabolomics analysis of GC–MS, LC-MS, CE-MS, and NMR-based chromatography.
Life 13 00706 g001
Figure 2. Environmental stresses of biotic and abiotic factors affecting the growth and homeostasis of plants.
Figure 2. Environmental stresses of biotic and abiotic factors affecting the growth and homeostasis of plants.
Life 13 00706 g002
Figure 3. Some eminent examples with medicinal properties of secondary plant metabolites are (A) salicin, (B) taxol (paclitaxel), and (C) morphine.
Figure 3. Some eminent examples with medicinal properties of secondary plant metabolites are (A) salicin, (B) taxol (paclitaxel), and (C) morphine.
Life 13 00706 g003
Figure 4. Chemical structures of some plants derived primery and scndary metabolites with key importance in different era of lfe. Among all, some commonly known alkaloids (A), cyanogenic glycosides (B), and (C) non-protein amino acids along with their protein amino acids analogues.
Figure 4. Chemical structures of some plants derived primery and scndary metabolites with key importance in different era of lfe. Among all, some commonly known alkaloids (A), cyanogenic glycosides (B), and (C) non-protein amino acids along with their protein amino acids analogues.
Life 13 00706 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Salam, U.; Ullah, S.; Tang, Z.-H.; Elateeq, A.A.; Khan, Y.; Khan, J.; Khan, A.; Ali, S. Plant Metabolomics: An Overview of the Role of Primary and Secondary Metabolites against Different Environmental Stress Factors. Life 2023, 13, 706. https://doi.org/10.3390/life13030706

AMA Style

Salam U, Ullah S, Tang Z-H, Elateeq AA, Khan Y, Khan J, Khan A, Ali S. Plant Metabolomics: An Overview of the Role of Primary and Secondary Metabolites against Different Environmental Stress Factors. Life. 2023; 13(3):706. https://doi.org/10.3390/life13030706

Chicago/Turabian Style

Salam, Uzma, Shakir Ullah, Zhong-Hua Tang, Ahmed A. Elateeq, Yaseen Khan, Jafar Khan, Asif Khan, and Sajid Ali. 2023. "Plant Metabolomics: An Overview of the Role of Primary and Secondary Metabolites against Different Environmental Stress Factors" Life 13, no. 3: 706. https://doi.org/10.3390/life13030706

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