Next Article in Journal
Effect of Composted Organic Waste on Miscanthus sinensis Andersson Yield, Morphological Characteristics and Chlorophyll Fluorescence and Content
Previous Article in Journal
Accelerated Iron Evolution in Quaternary Red Soils through Anthropogenic Land Use Activities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Lipidomics in Plants Under Abiotic Stress Conditions: An Overview

by
Juliane Maciel Henschel
1,
Antônio Nunes de Andrade
2,
Josefa Bruna Lima dos Santos
3,
Rodrigo Ribeiro da Silva
1,
Djair Alves da Mata
1,
Tancredo Souza
3 and
Diego Silva Batista
1,2,3,*
1
Programa de Pós-Graduação em Agronomia, Universidade Federal da Paraíba, Areia 58397-000, PB, Brazil
2
Departamento de Agricultura, Universidade Federal da Paraíba, Bananeiras 58220-000, PB, Brazil
3
Programa de Pós-Graduação em Ciências Agrárias (Agroecologia), Universidade Federal da Paraíba, Bananeiras 58220-000, PB, Brazil
*
Author to whom correspondence should be addressed.
Agronomy 2024, 14(8), 1670; https://doi.org/10.3390/agronomy14081670 (registering DOI)
Submission received: 12 July 2024 / Revised: 25 July 2024 / Accepted: 28 July 2024 / Published: 30 July 2024
(This article belongs to the Section Plant-Crop Biology and Biochemistry)

Abstract

:
Lipids are ubiquitous macromolecules that play essential roles in several metabolic processes in plants, such as primary and secondary metabolism, energy storage, and lipid signaling, also being major constituents of membranes. Considering their importance, lipid contents, proportion, and composition are widely modulated in response to environmental conditions, which is even more important under unfavorable conditions such as abiotic stresses. In recent years, technological advances have allowed for the analysis of the global lipid profile, also known as lipidomics, which has emerged as a powerful tool for the comprehensive analysis of the modulation and roles of lipids under different conditions. This review provides a current overview of plant lipidomics research, covering the different lipid classes found in plants, analytical techniques, and the main lipid-related responses under temperature, water, salt, alkali, heavy metal, nutrient deficiency, light, and oxidative stress.

1. Introduction

Cell membranes, including plasma and endomembranes, are the structures allowing for the compartmentalization that creates life. Membranes are composed of a variety of lipids that differ in their head group and fatty acid composition [1]. Membrane remodeling comprises the changes in the unsaturation degree and the proportion of different lipid classes, which alter membrane fluidity and permeability, and vesicle formation [2,3]. Among these lipids, the three main classes are glycerolipids, sphingolipids, and sterols [2].
Glycerolipids are the most abundant class of lipids, comprising phospholipids (PLs), galactolipids (GLs), triacylglycerols (TAGs), and sulfolipids (SLs) [4]. PLs, such as phosphatidylcholine (PC), phosphatidylethanolamine (PE), phosphatidylserine (PS), and phosphatidylinositol (PI), are the main components of biological membranes [5,6]. They consist of a glycerol backbone connecting two hydrophobic fatty acid chains with a hydrophilic phosphate head group, which can also contain choline, ethanolamine, serine, or inositol [2,6]. Due to their amphipathic nature, phospholipids play an essential role in maintaining the structure and function of cell membranes and organelles [7,8]. On the other hand, GLs, such as monogalactosyldiacylglycerol (MGDG) and digalactosyldiacylglycerol (DGDG), are the major components of thylakoid membranes [2,9]. MGDG and DGDG are also crucial for the assembly and function of the photosynthetic machinery [8,10]. Sphingolipids, in turn, comprise glycosyl inositolphosphoceramides (GIPCs), glucosylceramides (GCers), ceramides (Cers), and free long-chain bases (LCBs) [11]. This group of lipids contains a sphingosine backbone instead of a glycerol backbone and composes about 10% of total plant lipids [12,13]. Sphingolipids are components of the plasma membrane and endomembrane system, being crucial to the maintenance of membrane integrity [13,14]. The third group of membrane lipids are sterols, isoprenoid derivatives that are plant membrane components and precursors of hormones such as brassinosteroids [15,16]. The most abundant sterols in plants are β-sitosterol, campesterol, and stigmasterol [15]. Sterols are involved in maintaining membrane fluidity and permeability, cellular polarity, and they also play a role in signaling and stress response pathways [17].
Triacylglycerols (TAGs) are lipids composed of three fatty acids esterified to glycerol and are involved in energy storage, being the major component of vegetable oils [18,19]. Due to their high energy storage capacity, oils are widely used for food, feed, biodiesel, and industrial chemicals. In plants, TAGs are synthesized in the endoplasmic reticulum of most plant tissues, being stored in specialized organelles called lipid droplets (LDs) [20]; however, the major storage sites of TAGs in plants are seeds and fruits [21]. Under carbon and energy deficit conditions, TAGs are hydrolyzed into fatty acids and other metabolites to generate ATP and carbon skeletons by peroxisomal β-oxidation [20]. The breakdown of TAGs occurs primarily by lipases, also being reported to occur during autophagy in response to stresses [22]. Therefore, in addition to developmental processes such as seed germination, seedling establishment, stomatal opening, and flower development, TAGs are also important in stress responses [19,23,24]. For instance, TAG biosynthesis under stress conditions represents a strategy to store cytotoxic lipid intermediates generated by membrane lipid remodeling, such as free fatty acids, diacylglycerol (DAG), and other hydrolytic byproducts, which can lead to lipotoxicity [19,25]. In addition, LDs are required for the adequate function of LD-associated proteins such as caleosins, proteins involved in stress sensing and signaling, which are also involved in the biosynthesis of oxylipins [26]. As a result, biotic and abiotic stresses are often related to the increase in TAGs [19].
Autophagy is a catabolic pathway that degrades and recycles cellular components, also participating in developmental processes, stress responses, and programmed cell death [22,27]. Studies have reported that organellar autophagy is essential for lipid homeostasis, functioning in the lipid turnover via TAG synthesis [28]. Moreover, under carbon starvation conditions, LDs are degraded via lipophagy into fatty acids to produce energy through β-oxidation [28,29].
In addition to the barriers imposed by membranes, lipids can also create physical barriers on the epidermis surface. For instance, cutin and suberin are mainly composed of fatty acid (FA) derivatives, epoxyacids, polyhydroxyacids, and saturated very long-chain FAs, being major barriers of the plant epidermal surface and root endodermis [30,31,32]. Lipids are also involved in plant–microbe interactions, modulating the signaling, perception, or defenses with either beneficial or phytopathogenic microorganisms, functioning like chemical barriers against infection [33,34]. These signals can be emitted from plants to microbes, microbes to plants, or microbes to microbes and include rhizodeposits of FAs, sterols, lipopolysaccharides, peptidoglycans, flagellin, and chitin [33]. Several lipid classes, such as fatty acids, phospholipids, lysophospholipids, phosphatidic acid, DAGs, inositol phosphate, oxylipins, sphingolipids, and N–acylethanolamine, are also known to participate in plant stress signaling by inducing the transcription of defense-related genes and by functioning in membrane perception and trafficking processes [3,35,36,37].
Plants are exposed to a wide range of environmental stresses, including drought, heat, cold, salt, and herbivore and pathogen attacks. These stresses can significantly affect plant growth, development, and productivity, as well as their lipid regulation, synthesis, degradation, relocalization, or restructuring [38,39]. For instance, stress through high or low temperatures directly impacts membrane fluidity by altering the interactions and composition of membrane lipids [40], while biotic stresses can induce the production of defense lipids such as oils and waxes [41]. Moreover, biotic and abiotic stresses commonly lead to oxidative stress due to the overproduction of unstable free radicals, such as reactive oxygen species (ROS), reactive nitrogen species, reactive carbonyl species, and reactive sulfur species [42,43]. The excess of these free radicals causes oxidative damage to biomolecules such as lipids, altering membrane permeability and stability, disrupting electron transport chains, and finally leading to cell death processes, including autophagy [44]. In this context, lipidomics has emerged as a powerful tool to investigate and comprehend the lipid metabolism, including the consequences of stresses and, thus, providing information on how to overcome the negative effects of these stresses. Although many specific studies on lipidomics in plants have been conducted, a comprehensive overview of plant lipidomics, particularly concerning abiotic stress responses, has not been extensively reported on in the literature. To better understand the potential of lipidomics in the study of plant stress responses, we begin with an overview of the current status of lipidomics research. We then provide a brief review of the main lipid classes and the biological processes affected by lipid composition. Finally, we discuss the effects of different stressors on these lipid-related responses.

2. Overview of Current Research on Plant Lipidomics

Lipids are a diverse group of biomolecules that play important roles in a variety of biological processes, and recent advances in analytical techniques have led to the emergence of lipidomics [45,46]. This technique provides insights into the complex biological processes underlying plant growth, development, and adaptation to stresses [41,47]. Considering that abiotic and biotic stresses are major limiting factors in agriculture, the use of lipidomics is an essential tool in studying the role of lipids in plant stress responses.
Searches on the Scopus and PubMed databases indicate a growing number of publications containing the term “plant lipidomics” in their title, abstract, or keywords (Figure 1a). From the 773 publications about “plant lipidomics” found on the Scopus database, 388 were related to the subject area “biochemistry, genetics and molecular biology”, 351 were related to “agricultural and biological sciences”, and 181 were related to “chemistry” (Figure 1c). When limiting the search to “plant stress lipidomics”, 390 publications were found in the Scopus database, with 123 publications on “biochemistry, genetics, and molecular biology” and 122 publications on “agricultural and biological sciences” (Figure 1b,d). China was the country with the highest number of publications about “plant lipidomics” (304) and “plant stress lipidomics” (98), followed by the United States and Germany (Figure 1e,f). These numbers highlight the growing potential of lipidomics in the study of plant stress biology.
Up to now, the majority of publications on lipidomics and plant stresses have focused on abiotic stresses, with temperature (heat and cold stress) being the most studied environmental factor (Table 1). Most of these studies describe the overall changes in the lipid profile, linking them either to membrane- or signaling-related responses. The lipid profiles of genotypes with varying stress tolerance are often used to identify the underlying mechanism involved in stress responses [48,49,50]. Similarly, the lipidomics of plants treated with stress mitigators are also used to investigate those responses [51,52,53,54].
Mass spectrometry (MS), liquid chromatography (LC), and gas chromatography (GC) are the most commonly used techniques in plant lipidomics (Table 1). In addition, multiomic and integrated approaches, such as the combination of lipidomics, transcriptomics, and metabolomics, allow researchers to draw more comprehensive conclusions about the observed responses [52,53,77,124]. By combining these diverse datasets, multiomic analyses can reveal how stress conditions affect gene expression (genomics and transcriptomics), protein abundance and activity (proteomics), and metabolic profiles (metabolomics and lipidomics) simultaneously. This integrative approach allows for the identification of key regulatory networks, novel biomarkers, and stress-responsive pathways that might be overlooked when studying each omics layer in isolation.

3. Abiotic Stress Responses

In addition to their direct detrimental effects on plants, abiotic stresses also lead to secondary oxidative stress, with the overproduction of unstable molecules that oxidize unsaturated fatty acids, reducing membrane stability and increasing its fluidity and permeability [40]. As a result, plants change the levels of several PLs such as PC and PE to maintain membrane stability [3,40,125,126]. Considering the ubiquitous presence of oxidative stress during plant exposition to other stresses, its effects will be further discussed along with each stress.

3.1. Temperature Stress

In addition to oxidative stress, abiotic stresses also induce specific effects on plant membranes. For instance, heat stress induces the hyper-fluidity of membranes, protein denaturation, organelle swelling, and the separation of bilayer-forming lipids, dismantling membranes and producing cytotoxic carbonyl groups [40]. In response to these negative effects, plants adjust their membrane lipid composition by increasing the total phosphatidylglycerol, phosphatidylinositol, and saturated phosphatidylcholine and phosphatidylethanolamine [125]. Moreover, studies reported increases in galactolipids (linoleate), phospholipids (palmitate, stearate, and oleate), and triacylglycerols (α-linolenate and hexadecatrienoic acid) in chloroplasts, the endoplasmic reticulum, and the plasma membrane [40]. This increase in saturated and monounsaturated fatty acids (16:0, 18:0, and 18:1) results in higher membrane stiffness, counteracting the effects of heat stress, while the polyunsaturation of fatty acids is often related to lower heat tolerance [125]. In addition to membrane remodeling, lipids are essential for triggering signaling pathways and providing energy during stress responses. For instance, the catabolism of fatty acids is required for efficient stomatal opening during heat stress in Arabidopsis [90]. Regarding stress signaling, heat stress induces the accumulation of phosphatidylinositol through the activation of phosphatidylinositol phosphate kinase (PIPK), also inducing the accumulation of Ca2+, which activates calmodulin-binding kinases that, in turn, induce the expression of heat shock proteins [127].
Unlike heat stress, cold stress reduces membrane fluidity, and thus, tolerance responses comprise the increase in the contents of unsaturated lipids in relation to saturated lipids, increasing the fluidity of cell and plastid membranes [58,65,67,127]. Under cold stress, the influx of Ca2+ induces the accumulation of the signaling lipid phosphatidic acid, while increased MGDG is pointed to function in the synthesis of jasmonate, a phytohormone involved in cold stress tolerance [128]. In maize, cold stress tolerance is dependent on lipid remodeling by the modulation of the levels of several lipids, such as mono- and di-unsaturated lysophosphatidic acid, lysophosphatidylcholine, PC, and phosphatidylinositol, as well as polyunsaturated phosphatidic acid, MGDG, DAG, and TAG [68]. In cold-stressed maize roots, the major lipid classes are PC, phosphatidylethanolamine, phosphatidic acid, and phosphatidylinositol [67]. In barley, high levels of phosphatidic acid, lysophosphatidic acid, and MGDG are also reported to induce cold tolerance [60]. The cold-tolerant wild tomato Solanum habrochaites also shows higher contents of MGDG, DGDG, and phosphoinositide compared to cold-sensitive S. lycopersicum species, also showing the increased β-oxidation of fatty acids in mitochondria to produce energy under stress conditions [74]. Bryophytes living under low temperatures survive by producing very long-chain polyunsaturated fatty acids, accumulating phospholipids and glycolipids, and decreasing storage lipids [70]. Low temperatures are often used to increase the post-harvest life of several fruits; however, they can also cause chilling injuries and lipid remodeling [63]. In peach fruits, cold stress tolerance is associated with increased phosphatidylcholine and phosphatidylethanolamine (and DAGs and TAGS) and reduced phosphatidic acid accumulation, which reduces chilling injuries [129]. Blueberry, in turn, undergoes pitting during cold storage, which is followed by increases in DGDG, phosphatidic acid, phosphatidylserine, phosphatidylglycerol, phosphatidylinositol, lysophosphatidylcholine, and lysophosphatidylethanolamine and decreases in phosphatidylcholine [63].

3.2. Water Stress

Drought stress occurs when the water supply is below the plant requirements; it is a major stress impacting crop yield worldwide, mainly considering global climate changes [130,131]. The primary effect of drought in plants is cell dehydration, which triggers stomatal closure, oxidative stress, membrane and protein structure destabilization, and photosynthesis inhibition, ultimately leading to cell damage and death [131,132]. As a result, phospholipids such as phosphatidylcholine, phosphatidylethanolamine, and phosphatidylglycerol are the first to be degraded, dismantling membranes [127]. The production of signaling lipids, as well as lipid byproducts, is increased under drought, while essential oil production can be increased or decreased depending on the species [35,133,134,135]. The membrane’s peroxidation results in the increase in malondialdehyde, which is an important stress indicator [49,50]. Thus, to better resist drought, plants can modulate the synthesis and composition of several lipid classes. For instance, the increase in the cuticle deposition of cutin and waxes, as well as the composition of cuticle-associated very long-chain lipids, may seal the leaf surface, reducing transpiration, being an important tolerance response to drought stress [49,136]. For example, drought-tolerant wheat plants were found to increase the content of cutin, alkanoic acids, and 2-hydroxyacids while decreasing aldehydes, which was not observed in drought-sensitive plants [49]. The alkane pathway comprises the synthesis of aldehydes, secondary alcohols, and ketones, while the primary alcohol pathway results in the production of esters and primary alcohols present in waxes [137].
Membrane stability under drought, in turn, is increased through the synthesis of linolenic acid and the reduction in MGDG and fatty acid desaturation [75]. Photosynthetic capacity and membrane fluidity under drought are associated with higher total contents of DGDG and MGDG, and the DGDG/MGDG ratio, as they are the main components of photosynthetic membranes [50,79,127]. Other lipids modulated under drought stress include phosphatidylglycerol, phosphatidic acid, sulfoquinovosyldiacylglycerol, phosphatidylcholine, lyso-phosphatidylcholine, and phosphatidylinositol, pointing mainly to their function in membrane fluidity and lipid signaling [77]. Lipid signaling through phosphatidic acid, phosphatidyl serine, phosphatidyl glycerol, and phosphatidyl choline may also involve their interaction with dehydrins, which improve drought tolerance by increasing water retention capacity, chlorophyll content, antioxidant defense, and osmoregulation [138]. Moreover, these lipids, as well as sphingolipids and oxylipins, are known to interact with abscisic acid and nitric oxide, regulating stomatal closure [127,135]. In the opposite way, abscisic acid is also known to induce fatty acid biosynthesis [134]. In drought-tolerant thyme plants, the level of signaling lipids increases, while the level of oxylipins, which are induced by ROS, dramatically decreases [81]. The fact that the exogenous application of phosphatidylcholine increases drought tolerance in peach trees, increasing membrane stability, osmoprotection, and photosynthetic capacity, further supports the importance of these lipids under drought conditions [139].
Waterlogging or submergence stress is caused by the excess of water in the root zone in land plants, which leads to poor soil aeration and oxygen depletion to roots (hypoxia or anoxia), reducing aerobic respiration and energy production, inducing anaerobic fermentation, and ultimately impairing plant growth, development, and yield [140,141]. This stress induces the accumulation of ROS, leading to membrane peroxidation and electrolyte leakage, also reducing photosynthetic capacity due to damages to chloroplast membranes [85]. Hypoxia is reported to reduce the global contents of lipids, inhibiting their biosynthesis and inducing degradation, which culminates in free fatty acid accumulation [140]. Tolerance against waterlogging stress has been associated with lipid remodeling, with the changes being species- and stress duration-dependent. In ramie (Boehmeria nivea), for example, lipid droplet formation is increased in the inter-chloroplast cytoplasm, while fatty acid and 18:3 (n3) Coenzyme A (C18:3-CoA) contents are gradually increased [83]. Similarly, the unsaturation of very long-chain ceramides and acyl-CoAs is reported to reduce damages caused by hypoxia by interacting with ethylene pathways, while the modulation of wax deposition is also related to waterlogging adaptation [140,141,142]. Signaling lipids, such as phosphatidic acid, have been reported to increase under hypoxia [140]. In turn, lipids such as oxylipin jasmonate are important modulators of plant responses to reoxygenation, improving antioxidant capacity [141].

3.3. Salt and Alkali Stress

Salt stress is characterized as the accumulation of ions on the root zone, mostly sodium and chloride, which reduces the osmotic potential, impairing the uptake of water and nutrients [143,144,145]. Moreover, salt stress causes the overaccumulation of sodium on cells, disrupting electrochemical gradients and causing protein denaturation [146]. Alkali stress, also called saline–alkali stress, occurs when neutral or alkali salts accumulate in soils, increasing pH and causing chemical destruction, osmotic stress, ion toxicity, and nutrient and oxygen deficiency [147,148].
Membrane integrity and permeability are essential for the modulation of sodium uptake and compartmentalization [149]. Salt stress induces lipolysis and lipid peroxidation and reduces lipid biosynthesis, reducing total membrane lipid contents in salt-sensitive species; however, salt-tolerant species increase lipid homeostasis, also increasing total lipid content [150,151]. Salt stress also increases the saturation of fatty acids in sensitive species, resulting in more rigid membranes; however, halophytes show a high proportion of unsaturated fatty acids, maintaining membrane fluidity [151]. In addition, salt-tolerant species increase the contents and proportion of specific lipid groups, such as phospholipids [126]. For instance, the content of phosphatidylcholine, phosphatidylethanolamine, phosphatidylinositol, and phosphatidylserine are increased in these species, with high phosphatidylcholine-to-phosphatidylethanolamine (PC/PE) ratios also being related to increased salt tolerance [126,150,152,153]. The accumulation of glycerophosphocholine has shown osmolyte properties, also being a substrate for PC synthesis in barley roots [76]. Moreover, the increase in PLs is also observed in the membrane of vacuoles, highlighting the role of these lipids in increasing membrane stability [122]. Increases in phosphatidylinositol and phosphatidylserine were also observed under salt stress, being implicated in signaling responses through Ca2+ that trigger Na+ efflux [122,123,154,155]. In chloroplast membranes, the increase in phosphatidylglycerol is related to higher membrane stability, maintaining the function of light-harvesting complexes and photosystem II [156,157].
The decrease in glycolipid content, such as MGDG and DGDG, is related to the negative effects of salinity in sensitive species [158], while its increase is related to the higher stability of the photosynthetic apparatus [152]. The proportion between these two glycolipids is essential to confer tolerance against salt stress, as the increase in the DGDG/MGDG ratio stabilizes the membrane by increasing the ratio of bilayer-forming to non-bilayer-forming, while salinity often reduces the DGDG/MGDG ratio in sensitive plants [156,159,160]. The breakdown of DGDG and MGDG during salt stress generates polyunsaturated fatty acids that are involved in jasmonate biosynthesis, increasing the levels of this hormone [159]. Sulfoquinovosyl diacylglycerol (SQDG) is another glycolipid associated with the lipid–protein interactions of the photosynthetic apparatus, whose increase is often related to salt tolerance [161]. Sphingolipids, such as PhytoSph, PhytoCer, PhytoCer-OHFA, Phyto-GluCer, and phosphatidyl ceramide, are also reported to alter the tolerance against salt stress in Arabidopsis, with their proportion to other lipid groups being essential for tolerance [120].
Similar to the other lipid groups, increased sterol contents are related to higher salt tolerance, while the proportion between sterol groups in response to salinity varies according to the tissue [162]. Moreover, the proportion between sterols and phospholipids is often increased in salt-tolerant species, resulting in lower membrane permeability to Na+ and Cl [163]. LD accumulation is also known to improve recovery after salt stress, probably by providing fatty acids used in membrane reconstruction [164].
In addition to salinity, alkali stress is also known to induce deep lipid remodeling. In sunflower, for example, alkali stress increases the contents of saccharolipids and glycerolipids and reduces the contents of glycerophospholipids in sunflower seeds, also affecting seed oil quality [56]. Maize roots also show extensive lipid remodeling under alkali stress, with reductions in the contents of phosphatidylcholine and inductions in the synthesis of phospholipids and galactolipids in plastids and the production of C34:6 galactolipids, pointing to the induction of the prokaryotic pathway [55]. The modulation of root exudates is also an important defense response against alkali stress. In wheat, alkali stress was reported to increase glycolysis, fatty acid synthesis, and phenolic acid synthesis [165]. In the halophyte Leymus chinensis, the accumulation and secretion of organic acids, amino acids, and fatty acids are involved in alkali tolerance [166].

3.4. Heavy Metals

Heavy metal stress induces significant changes in plant metabolism as a means of coping with stress or even eliminating heavy metals. One of the primary changes observed is lipid peroxidation, which can compromise the integrity of the plasma membrane [167,168]. Additionally, heavy metal stress can lead to both qualitative and quantitative changes in membrane lipids, altering the membrane’s permeability and function [169]. For instance, the presence of mercury (Hg) in the environment alters the lipid pattern in the leaves of the saltmarsh halophyte Halimione portulacoides. This alteration affects the level of unsaturation, fluidity, and stability of the membrane, which in turn influences the plant’s tolerance to Hg [97].
Cadmium (Cd) is another heavy metal that significantly impacts the composition of lipophilic compounds in both leaves and roots. Changes in lipid composition under sublethal Cd toxicity likely serve as adjustment mechanisms to maintain membrane fluidity [98]. In wheat roots, Cd stress results in decreased concentrations of several key lipids, including digalactosyldiacylglycerol, monogalactosyldiacylglycerol, phosphatidylcholine, phosphatidylethanolamine, phosphatidylglycerol, phosphatidylserine, and lysophosphatidylcholine. Additionally, the concentrations of specific glycerolipids such as C36:4, C34:2, C34:3, and C36:6 are also reduced under Cd stress [99].

3.5. Nutrient Deficiency

When the availability of any of the 14 essential mineral nutrients is below the plant requirements, it is called nutritional deficiency, which results in reduced growth, production, and even death [170]. As these nutrients have unique functions in plants, the effects of their deficiencies will vary according to the nutrient, its levels, source, and plant species [170,171]. For instance, nitrogen is a major component of proteins and macromolecules, such as RuBisCo and chlorophylls, and thus, its deficiency affects overall plant metabolism, also harming photosynthetic machinery and capacity [171,172]. Phosphorus, in turn, is involved in energy metabolism, nucleic acids, and phospholipids, the main components of membranes, while potassium is an important osmolyte regulating cell turgor and stomatal opening [172,173].
Nitrogen depletion is related to decreases in total lipid contents, while higher tolerance is associated with higher fatty acid unsaturation and chloroplast membrane stability [113,116]. Moreover, nitrogen deficiency is related to lower contents of phosphatidylglycerol and glucuronosyl diacylglycerol, as well as a lower MGDG/DGDG ratio [111]. In oleaginous microalga species, nitrogen and phosphorus starvation differentially affect the production and composition of storage lipids, with nutrient deficiency resulting in a higher production of triacylglycerols enriched in important commercial biochemicals, such as the long-chain polyunsaturated fatty acid arachidonic acid [107,110]. Similarly, the accumulation of triacylglycerols is also affected by nitrogen depletion in tea plants [114]. In soybean, nitrogen deficiency mainly affects the content of galactolipids, compared to phospholipids, with nodulated roots accumulating more phospholipids and showing a higher MGDG/DGDG ratio [113].
Phosphorus deficiency causes extensive membrane lipid remodeling, including the change of phospholipids to non-phospholipids, with the ability to maintain the content of phospholipids being an important tolerance response [112,174]. As a consequence of lower phospholipid contents, plants increase the contents of membrane galactolipids, also increasing triacylglycerol contents in leaves and roots [115]. In Arabidopsis and rice, in addition to membrane remodeling, the presence of glucuronosyldiacylglycerol is related to higher tolerance to phosphorus depletion [106].

3.6. Light

Light quantity, quality, and photoperiod are shown to affect lipid dynamics and composition in plants [175,176]. High irradiances induce higher photosynthetic rates and consequently, a greater accumulation of lipids. Additionally, there are changes in lipid composition under high light conditions [175]. Blue light, in particular, stimulates lipid biosynthesis in plants. This has been demonstrated in tea plants, where high levels of blue light (200 μmol m–2 s–1) upregulated 15 genes involved in lipid synthesis [104]. Blue light increases lipid biosynthesis by inducing the flux of fatty acid biosynthesis, the Calvin cycle, and the Krebs cycle [177]. In contrast, Arabidopsis thaliana plants grown under amber light (595 nm) also show a greater accumulation of lipids, likely linked to oxidative stress. The most pronounced increases were in tocopherols, which play a crucial role in eliminating singlet oxygen [100,175]. Tocopherols are also responsible for initiating stress responses caused by high light irradiance in tomatoes [105].
Increased solar radiation can lead to higher levels of fatty acids stored in seeds and alter their composition, resulting in a greater proportion of C18:1 fatty acids and lower amounts of C18:2 fatty acids [102,175]. Analyzing the response of membrane lipid composition in plants under different light intensities reveals that light significantly affects lipid levels. Darkness induces a decrease in glycerolipids, whereas high light conditions lead to an increase in these lipids [101].

4. Concluding Remarks

Great variation in the lipidomic results was observed among species and abiotic stresses, with most conclusions pointing to membrane remodeling and lipid signaling. Few studies, however, report on the mechanisms involved in energy metabolism and lipid catabolism and anabolism. Also, studies about crosstalk among abiotic stressors or that relate the role of lipids and plant hormones in the response to abiotic stressors are missing. In this context, most efforts in plant lipidomics have been directed towards exploratory changes in the contents, proportions, and compositions of lipids. However, deeper conclusions can be drawn by using integrative approaches, which allow for a comprehensive analysis not only of lipid composition but also of the underlying mechanisms involved in the responses. Another promising field is the identification of new lipids with abiotic stress mitigation functions. This could enable the development of new products and tolerant genotypes, as well as the manipulation of environmental conditions to produce specific lipids with commercial appeal. Moreover, as a currently evolving field, omics approaches, including lipidomics, have great potential to help researchers elucidate plant responses to abiotic stresses, which is paramount under the current climate change scenario.

Author Contributions

Conceptualization, J.M.H. and D.S.B.; methodology, J.M.H. and D.S.B.; investigation, J.M.H. and D.S.B.; data curation, J.M.H. and D.S.B.; writing—original draft preparation, J.M.H., R.R.d.S., D.A.d.M., A.N.d.A., J.B.L.d.S., T.S. and D.S.B.; writing—review and editing, J.M.H., T.S. and D.S.B.; visualization, J.M.H., T.S. and D.S.B.; supervision, D.S.B.; project administration, D.S.B.; funding acquisition, J.M.H. and D.S.B. All authors have read and agreed to the published version of the manuscript.

Funding

This study was partially funded by the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq, Brasília, DF, Brazil: Grants no. 301858/2023-3 to J.M.H. and 304214/2022-1 to D.S.B.) and the Research Support Foundation of the State of Paraíba (FAPESQ: Post-doc Research Fellowship to T.S. and Grant “EDITAL-006/2020-PDCTR-PB-2020” to J.M.H.).

Data Availability Statement

Data will be made available on reasonable request to the corresponding author.

Acknowledgments

We thank the research group in Plant Physiology at the Federal University of Paraíba (GFV/UFPB) for the insights that generated the idea for this review. We also thank the National Council for Scientific and Technological Development (CNPq—Brazil), the Research Support Foundation of the State of Paraíba (FAPESQ), the Federal University of Paraíba (UFPB), and Coordination for the Improvement of Higher Education Personnel (CAPES) for the scholarships granted to the researchers.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Cassim, A.M.; Gouguet, P.; Gronnier, J.; Laurent, N.; Germain, V.; Grison, M.; Boutté, Y.; Gerbeau-Pissot, P.; Simon-Plas, F.; Mongrand, S. Plant Lipids: Key Players of Plasma Membrane Organization and Function. Prog. Lipid Res. 2019, 73, 1–27. [Google Scholar] [CrossRef] [PubMed]
  2. Reszczyńska, E.; Hanaka, A. Lipids Composition in Plant Membranes. Cell Biochem. Biophys. 2020, 78, 401–414. [Google Scholar] [CrossRef] [PubMed]
  3. Xing, J.; Zhang, L.; Duan, Z.; Lin, J. Coordination of Phospholipid-Based Signaling and Membrane Trafficking in Plant Immunity. Trends Plant Sci. 2021, 26, 407–420. [Google Scholar] [CrossRef] [PubMed]
  4. MacDonald, G.E.; Lada, R.R.; Caldwell, C.D.; Udenigwe, C.; MacDonald, M.T. Potential Roles of Fatty Acids and Lipids in Postharvest Needle Abscission Physiology. Am. J. Plant Sci. 2019, 10, 1069–1089. [Google Scholar] [CrossRef]
  5. Nakamura, Y. Plant Phospholipid Diversity: Emerging Functions in Metabolism and Protein–Lipid Interactions. Trends Plant Sci. 2017, 22, 1027–1040. [Google Scholar] [CrossRef]
  6. Colin, L.A.; Jaillais, Y. Phospholipids across Scales: Lipid Patterns and Plant Development. Curr. Opin. Plant Biol. 2020, 53, 1–9. [Google Scholar] [CrossRef] [PubMed]
  7. Lin, F.; Qu, Y.; Zhang, Q. Phospholipids: Molecules Regulating Cytoskeletal Organization in Plant Abiotic Stress Tolerance. Plant Signal. Behav. 2014, 9, e28337. [Google Scholar] [CrossRef] [PubMed]
  8. Hölzl, G.; Dörmann, P. Chloroplast Lipids and Their Biosynthesis. Annu. Rev. Plant Biol. 2019, 70, 51–81. [Google Scholar] [CrossRef] [PubMed]
  9. Kalisch, B.; Dörmann, P.; Hölzl, G. DGDG and Glycolipids in Plants and Algae. Subcell. Biochem. 2016, 86, 51–83. [Google Scholar] [CrossRef] [PubMed]
  10. Basnet, R.; Zhang, J.; Hussain, N.; Shu, Q. Characterization and Mutational Analysis of a Monogalactosyldiacylglycerol Synthase Gene OsMGD2 in Rice. Front. Plant Sci. 2019, 10, 992. [Google Scholar] [CrossRef] [PubMed]
  11. Sato, M.; Nagano, M.; Jin, S.; Miyagi, A.; Yamaguchi, M.; Kawai-Yamada, M.; Ishikawa, T. Plant-Unique Cis/Trans Isomerism of Long-Chain Base Unsaturation Is Selectively Required for Aluminum Tolerance Resulting from Glucosylceramide-Dependent Plasma Membrane Fluidity. Plants 2020, 9, 19. [Google Scholar] [CrossRef] [PubMed]
  12. Dunn, T.; Lynch, D.; Michaelson, L.; Napier, J. A Post-genomic Approach to Understanding Sphingolipid Metabolism in Arabidopsis thaliana. Ann. Bot. 2004, 93, 483–497. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, N.J.; Hou, L.P.; Bao, J.J.; Wang, L.J.; Chen, X.Y. Sphingolipid Metabolism, Transport, and Functions in Plants: Recent Progress and Future Perspectives. Plant Commun. 2021, 2, 100214. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, S.; Chen, M.; Li, Y.; Huang, X.; Niu, D.; Rashid, A.; Xu, C.; Wang, K. Adjustments of Both Phospholipids and Sphingolipids Contribute to Cold Tolerance in Stony Hard Peach Fruit by Continuous Ethylene. Postharvest Biol. Technol. 2021, 171, 111332. [Google Scholar] [CrossRef]
  15. Valitova, J.N.; Sulkarnayeva, A.G.; Minibayeva, F.V. Plant Sterols: Diversity, Biosynthesis, and Physiological Functions. Biochemistry 2016, 81, 819–834. [Google Scholar] [CrossRef] [PubMed]
  16. Tarkowská, D.; Krampolová, E.; Strnad, M. Plant Triterpenoid Crosstalk: The Interaction of Brassinosteroids and Phytoecdysteroids in Lepidium sativum. Plants 2020, 9, 1325. [Google Scholar] [CrossRef] [PubMed]
  17. Kato, H.; Nemoto, K.; Shimizu, M.; Abe, A.; Asai, S.; Ishihama, N.; Matsuoka, S.; Daimon, T.; Ojika, M.; Kawakita, K.; et al. Recognition of Pathogen-Derived Sphingolipids in Arabidopsis. Science 2022, 376, 857–860. [Google Scholar] [CrossRef] [PubMed]
  18. Bates, P.D. Understanding the Control of Acyl Flux through the Lipid Metabolic Network of Plant Oil Biosynthesis. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2016, 1861, 1214–1225. [Google Scholar] [CrossRef]
  19. Lu, J.; Xu, Y.; Wang, J.; Singer, S.D.; Chen, G. The Role of Triacylglycerol in Plant Stress Response. Plants 2020, 9, 472. [Google Scholar] [CrossRef] [PubMed]
  20. Xu, C.; Shanklin, J. Triacylglycerol Metabolism, Function, and Accumulation in Plant Vegetative Tissues. Annu. Rev. Plant Biol. 2016, 67, 179–206. [Google Scholar] [CrossRef] [PubMed]
  21. Chapman, K.D.; Aziz, M.; Dyer, J.M.; Mullen, R.T. Mechanisms of Lipid Droplet Biogenesis. Biochem. J. 2019, 476, 1929–1942. [Google Scholar] [CrossRef] [PubMed]
  22. Barros, J.A.S.; Siqueira, J.A.B.; Cavalcanti, J.H.F.; Araújo, W.L.; Avin-Wittenberg, T. Multifaceted Roles of Plant Autophagy in Lipid and Energy Metabolism. Trends Plant Sci. 2020, 25, 1141–1153. [Google Scholar] [CrossRef] [PubMed]
  23. Yang, Y.; Benning, C. Functions of Triacylglycerols during Plant Development and Stress. Curr. Opin. Biotechnol. 2018, 49, 191–198. [Google Scholar] [CrossRef]
  24. Wang, L.; Qian, B.; Zhao, L.; Liang, M.H.; Zhan, X.; Zhu, J. Two Triacylglycerol Lipases Are Negative Regulators of Chilling Stress Tolerance in Arabidopsis. Int. J. Mol. Sci. 2022, 23, 3380. [Google Scholar] [CrossRef] [PubMed]
  25. Yu, L.; Zhou, C.; Fan, J.; Shanklin, J.; Xu, C. Mechanisms and Functions of Membrane Lipid Remodeling in Plants. Plant J. 2021, 107, 37–53. [Google Scholar] [CrossRef] [PubMed]
  26. Ischebeck, T.; Krawczyk, H.E.; Mullen, R.T.; Dyer, J.M.; Chapman, K.D. Lipid Droplets in Plants and Algae: Distribution, Formation, Turnover and Function. Semin. Cell Dev. Biol. 2020, 108, 82–93. [Google Scholar] [CrossRef] [PubMed]
  27. Wang, P.; Wang, T.; Han, J.; Li, M.; Zhao, Y.; Su, T.; Ma, C. Plant Autophagy: An Intricate Process Controlled by Various Signaling Pathways. Front. Plant Sci. 2021, 12, 754982. [Google Scholar] [CrossRef] [PubMed]
  28. Farquharson, K.L. Autophagy Contributes to Plant Lipid Homeostasis. Plant Cell 2019, 31, 1427–1428. [Google Scholar] [CrossRef] [PubMed]
  29. Barros, J.A.S.; Magen, S.; Lapidot-Cohen, T.; Rosental, L.; Brotman, Y.; Araújo, W.L.; Avin-Wittenberg, T. Autophagy Is Required for Lipid Homeostasis during Dark-Induced Senescence. Plant Physiol. 2021, 185, 1547–1558. [Google Scholar] [CrossRef] [PubMed]
  30. Cahoon, E.B.; Li-Beisson, Y. Plant Unusual Fatty Acids: Learning from the Less Common. Curr. Opin. Plant Biol. 2020, 55, 66–73. [Google Scholar] [CrossRef] [PubMed]
  31. He, M.; Ding, N.Z. Plant Unsaturated Fatty Acids: Multiple Roles in Stress Response. Front. Plant Sci. 2020, 11, 562785. [Google Scholar] [CrossRef] [PubMed]
  32. Philippe, G.; Sørensen, I.; Jiao, C.; Sun, X.; Fei, Z.; Domozych, D.S.; Rose, J.K. Cutin and Suberin: Assembly and Origins of Specialized Lipidic Cell Wall Scaffolds. Curr. Opin. Plant Biol. 2020, 55, 11–20. [Google Scholar] [CrossRef] [PubMed]
  33. Macabuhay, A.; Arsova, B.; Walker, R.; Johnson, A.; Watt, M.; Roessner, U. Modulators or Facilitators? Roles of Lipids in Plant Root–Microbe Interactions. Trends Plant Sci. 2022, 27, 180–190. [Google Scholar] [CrossRef] [PubMed]
  34. Cavaco, A.R.; Matos, A.R.; Figueiredo, A. Speaking the Language of Lipids: The Cross-Talk between Plants and Pathogens in Defence and Disease. Cell. Mol. Life Sci. 2021, 78, 4399–4415. [Google Scholar] [CrossRef] [PubMed]
  35. 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] [PubMed]
  36. Huby, E.; Napier, J.A.; Baillieul, F.; Michaelson, L.V.; Dhondt-Cordelier, S. Sphingolipids: Towards an Integrated View of Metabolism during the Plant Stress Response. New Phytol. 2020, 225, 659–670. [Google Scholar] [CrossRef] [PubMed]
  37. Wang, J.; Chen, Y.L.; Li, Y.K.; Chen, D.K.; He, J.F.; Yao, N. Functions of Sphingolipids in Pathogenesis During Host–Pathogen Interactions. Front. Microbiol. 2021, 12, 701041. [Google Scholar] [CrossRef] [PubMed]
  38. Koelmel, J.P.; Napolitano, M.P.; Ulmer, C.Z.; Vasiliou, V.; Garrett, T.J.; Yost, R.A.; Prasad, M.N.V.; Godri Pollitt, K.J.; Bowden, J.A. Environmental Lipidomics: Understanding the Response of Organisms and Ecosystems to a Changing World. Metabolomics 2020, 16, 56. [Google Scholar] [CrossRef] [PubMed]
  39. Waadt, R.; Seller, C.A.; Hsu, P.-K.; Takahashi, Y.; Munemasa, S.; Schroeder, J.I. Plant Hormone Regulation of Abiotic Stress Responses. Nat. Rev. Mol. Cell Biol. 2022, 23, 680–694. [Google Scholar] [CrossRef] [PubMed]
  40. Higashi, Y.; Saito, K. Lipidomic Studies of Membrane Glycerolipids in Plant Leaves under Heat Stress. Prog. Lipid Res. 2019, 75, 100990. [Google Scholar] [CrossRef] [PubMed]
  41. Tenenboim, H.; Burgos, A.; Willmitzer, L.; Brotman, Y. Using Lipidomics for Expanding the Knowledge on Lipid Metabolism in Plants. Biochimie 2016, 130, 91–96. [Google Scholar] [CrossRef] [PubMed]
  42. Farooq, M.A.; Niazi, A.K.; Akhtar, J.; Saifullah; Farooq, M.; Souri, Z.; Karimi, N.; Rengel, Z. Acquiring Control: The Evolution of ROS-Induced Oxidative Stress and Redox Signaling Pathways in Plant Stress Responses. Plant Physiol. Biochem. 2019, 141, 353–369. [Google Scholar] [CrossRef] [PubMed]
  43. Zhou, X.; Joshi, S.; Patil, S.; Khare, T.; Kumar, V. Reactive Oxygen, Nitrogen, Carbonyl and Sulfur Species and Their Roles in Plant Abiotic Stress Responses and Tolerance. J. Plant Growth Regul. 2022, 41, 119–142. [Google Scholar] [CrossRef]
  44. Signorelli, S.; Tarkowski, Ł.P.; Van den Ende, W.; Bassham, D.C. Linking Autophagy to Abiotic and Biotic Stress Responses. Trends Plant Sci. 2019, 24, 413–430. [Google Scholar] [CrossRef] [PubMed]
  45. Gross, R.W. The Evolution of Lipidomics through Space and Time. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2017, 1862, 731–739. [Google Scholar] [CrossRef] [PubMed]
  46. Han, X.; Gross, R.W. The Foundations and Development of Lipidomics. J. Lipid Res. 2022, 63, 100164. [Google Scholar] [CrossRef] [PubMed]
  47. Kehelpannala, C.; Rupasinghe, T.; Hennessy, T.; Bradley, D.; Ebert, B.; Roessner, U. The State of the Art in Plant Lipidomics. Mol. Omics 2021, 17, 894–910. [Google Scholar] [CrossRef]
  48. Narayanan, S.; Zoong-Lwe, Z.S.; Gandhi, N.; Welti, R.; Fallen, B.; Smith, J.R.; Rustgi, S. Comparative Lipidomic Analysis Reveals Heat Stress Responses of Two Soybean Genotypes Differing in Temperature Sensitivity. Plants 2020, 9, 457. [Google Scholar] [CrossRef]
  49. Zhang, X.; Ni, Y.; Xu, D.; Busta, L.; Xiao, Y.; Jetter, R.; Guo, Y. Integrative Analysis of the Cuticular Lipidome and Transcriptome of Sorghum Bicolor Reveals Cultivar Differences in Drought Tolerance. Plant Physiol. Biochem. 2021, 163, 285–295. [Google Scholar] [CrossRef] [PubMed]
  50. Hassan, M.J.; Qi, H.; Cheng, B.; Hussain, S.; Peng, Y.; Liu, W.; Feng, G.; Zhao, J.; Li, Z. Enhanced Adaptability to Limited Water Supply Regulated by Diethyl Aminoethyl Hexanoate (DA-6) Associated With Lipidomic Reprogramming in Two White Clover Genotypes. Front. Plant Sci. 2022, 13, 879331. [Google Scholar] [CrossRef] [PubMed]
  51. Hu, Q.; Cui, H.; Ma, C.; Li, Y.; Yang, C.; Wang, K.; Sun, Y. Lipidomic Metabolism Associated with Acetic Acid Priming-Induced Salt Tolerance in Carex rigescens. Plant Physiol. Biochem. 2021, 167, 665–677. [Google Scholar] [CrossRef] [PubMed]
  52. Buffagni, V.; Zhang, L.; Senizza, B.; Rocchetti, G.; Ferrarini, A.; Miras-Moreno, B.; Lucini, L. Metabolomics and Lipidomics Insight into the Effect of Different Polyamines on Tomato Plants under Non-Stress and Salinity Conditions. Plant Sci. 2022, 322, 111346. [Google Scholar] [CrossRef] [PubMed]
  53. Li, Z.; Cheng, B.; Zhao, Y.; Luo, L.; Zhang, Y.; Feng, G.; Han, L.; Peng, Y.; Zhang, X. Metabolic Regulation and Lipidomic Remodeling in Relation to Spermidine-Induced Stress Tolerance to High Temperature in Plants. Int. J. Mol. Sci. 2022, 23, 12247. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, X.; Gao, Y.; Zhuang, L.; Huang, B. Phosphatidic Acid Priming-Enhanced Heat Tolerance in Tall Fescue (Festuca arundinacea) Involves Lipidomic Reprogramming of Lipids for Membrane Stability and Stress Signaling. Plant Growth Regul. 2023, 99, 527–538. [Google Scholar] [CrossRef]
  55. Xu, X.; Zhang, J.; Yan, B.; Wei, Y.; Ge, S.; Li, J.; Han, Y.; Li, Z.; Zhao, C.; Xu, J. The Adjustment of Membrane Lipid Metabolism Pathways in Maize Roots Under Saline–Alkaline Stress. Front Plant Sci. 2021, 12, 635327. [Google Scholar] [CrossRef] [PubMed]
  56. Lu, H.; Wang, Z.; Xu, C.; Li, L.; Yang, C. Multiomics Analysis Provides Insights into Alkali Stress Tolerance of Sunflower (Helianthus annuus L.). Plant Physiol. Biochem. 2021, 166, 66–77. [Google Scholar] [CrossRef] [PubMed]
  57. Tarazona, P.; Feussner, K.; Feussner, I. An Enhanced Plant Lipidomics Method Based on Multiplexed Liquid Chromatography-Mass Spectrometry Reveals Additional Insights into Cold- and Drought-Induced Membrane Remodeling. Plant J. 2015, 84, 621–633. [Google Scholar] [CrossRef] [PubMed]
  58. Chen, M.; Thelen, J.J. Acyl-Lipid Desaturase 1 Primes Cold Acclimation Response in Arabidopsis. Physiol. Plant 2016, 158, 11–22. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, J.; Li, Q.; Chen, J.; Jiang, Y. Revealing Further Insights on Chilling Injury of Postharvest Bananas by Untargeted Lipidomics. Foods 2020, 9, 894. [Google Scholar] [CrossRef] [PubMed]
  60. Zhao, Y.; Li, S.; Wu, J.; Liu, H.; Wang, P.; Xu, L. Insights into Membrane Lipids Modification in Barley Leaves as an Adaptation Mechanism to Cold Stress. Plant Growth Regul. 2024, 103, 369–388. [Google Scholar] [CrossRef]
  61. Kong, X.; Wei, B.; Gao, Z.; Zhou, Y.; Shi, F.; Zhou, X.; Zhou, Q.; Ji, S. Changes in Membrane Lipid Composition and Function Accompanying Chilling Injury in Bell Peppers. Plant Cell Physiol. 2018, 59, 167–178. [Google Scholar] [CrossRef] [PubMed]
  62. Xu, D.; Lam, S.M.; Zuo, J.; Yuan, S.; Lv, J.; Shi, J.; Gao, L.; Chen, B.; Sui, Y.; Shui, G.; et al. Lipidomics Reveals the Difference of Membrane Lipid Catabolism between Chilling Injury Sensitive and Non-Sensitive Green Bell Pepper in Response to Chilling. Postharvest Biol. Technol. 2021, 182, 111714. [Google Scholar] [CrossRef]
  63. Wang, Y.; Ji, S.; Dai, H.; Kong, X.; Hao, J.; Wang, S.; Zhou, X.; Zhao, Y.; Wei, B.; Cheng, S.; et al. Changes in Membrane Lipid Metabolism Accompany Pitting in Blueberry during Refrigeration and Subsequent Storage at Room Temperature. Front. Plant Sci. 2019, 10, 829. [Google Scholar] [CrossRef] [PubMed]
  64. Zainal, P.W.; Syukri, D.; Fahmy, K.; Imaizumi, T.; Thammawong, M.; Tsuta, M.; Nagata, M.; Nakano, K. Lipidomic Profiling to Assess the Freshness of Stored Cabbage. Food Anal. Methods 2023, 16, 304–317. [Google Scholar] [CrossRef]
  65. Dhaliwal, L.K.; Shim, J.; Auld, D.; Angeles-Shim, R.B. Fatty Acid Unsaturation Improves Germination of Upland Cotton (Gossypium Hirsutum) under Cold Stress. Front. Plant Sci. 2024, 15, 1286908. [Google Scholar] [CrossRef] [PubMed]
  66. Zhan, X.; Qian, Y.; Mao, B. Dendrobium Multi-Omics Reveal Lipid Remodeling in Response to Freezing. Metabolites 2022, 12, 1216. [Google Scholar] [CrossRef] [PubMed]
  67. Zhao, X.; Wei, Y.; Zhang, J.; Yang, L.; Liu, X.; Zhang, H.; Shao, W.; He, L.; Li, Z.; Zhang, Y.; et al. Membrane Lipids’ Metabolism and Transcriptional Regulation in Maize Roots Under Cold Stress. Front. Plant Sci. 2021, 12, 639132. [Google Scholar] [CrossRef] [PubMed]
  68. Gao, L.; Jiang, H.; Li, M.; Wang, D.; Xiang, H.; Zeng, R.; Chen, L.; Zhang, X.; Zuo, J.; Yang, S.; et al. Genetic and Lipidomic Analyses Reveal the Key Role of Lipid Metabolism for Cold Tolerance in Maize. J. Genet. Genom. 2024, 51, 326–337. [Google Scholar] [CrossRef] [PubMed]
  69. Zheng, G.; Tian, B.; Li, W. Membrane Lipid Remodelling of Meconopsis Racemosa after Its Introduction into Lowlands from an Alpine Environment. PLoS ONE 2014, 9, e106614. [Google Scholar] [CrossRef] [PubMed]
  70. Lu, Y.; Eiriksson, F.F.; Thorsteinsdóttir, M.; Simonsen, H.T. Lipidomic Analysis of Moss Species Bryum Pseudotriquetrum and Physcomitrium Patens under Cold Stress. Plant Environ. Interac. 2022, 3, 254–263. [Google Scholar] [CrossRef] [PubMed]
  71. Liu, T.; Chen, J.; Xu, F.; He, X.; Yang, S.; Zhu, Y.; Li, W.; Zheng, G. Analysis of Changes in the Panax Notoginseng Glycerolipidome in Response to Long-Term Chilling and Heat. Plant Divers. 2020, 42, 102–110. [Google Scholar] [CrossRef] [PubMed]
  72. Zhu, Y.; Wang, K.; Wu, C.; Zhao, Y.; Yin, X.; Zhang, B.; Grierson, D.; Chen, K.; Xu, C. Effect of Ethylene on Cell Wall and Lipid Metabolism during Alleviation of Postharvest Chilling Injury in Peach. Cells 2019, 8, 1612. [Google Scholar] [CrossRef] [PubMed]
  73. Cheong, B.E.; Ho, W.W.H.; Biddulph, B.; Wallace, X.; Rathjen, T.; Rupasinghe, T.W.T.; Roessner, U.; Dolferus, R. Phenotyping Reproductive Stage Chilling and Frost Tolerance in Wheat Using Targeted Metabolome and Lipidome Profiling. Metabolomics 2019, 15, 144. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, K.; Zhao, X.; Bi, M.; Zhang, H.; He, Y.; Cui, Y.; Qi, M. Membrane Lipid Metabolism and Scavenging of Reactive Oxygen Species Are Essential for Cold Resistance of Solanum habrochaites (LA1777): Lipidomics and Transcriptomes Analysis. Environ. Exp. Bot. 2023, 209, 105290. [Google Scholar] [CrossRef]
  75. Rabeh, K.; Sbabou, L.; Rachidi, F.; Ferradouss, A.; Laghmari, G.; Aasfar, A.; El Arroussi, H.; Ouajdi, M.; El Antry, S.; Belkadi, B.; et al. Lipidomic Profiling of Argania spinosa L. (Skeels) Following Drought Stress. Appl. Biochem. Biotechnol. 2023, 195, 1781–1799. [Google Scholar] [CrossRef] [PubMed]
  76. Sarabia, L.D.; Boughton, B.A.; Rupasinghe, T.; Callahan, D.L.; Hill, C.B.; Roessner, U. Comparative Spatial Lipidomics Analysis Reveals Cellular Lipid Remodelling in Different Developmental Zones of Barley Roots in Response to Salinity. Plant Cell Environ. 2020, 43, 327–343. [Google Scholar] [CrossRef] [PubMed]
  77. Shen, J.; Wang, S.; Sun, L.; Wang, Y.; Fan, K.; Li, C.; Wang, H.; Bi, C.; Zhang, F.; Ding, Z. Dynamic Changes in Metabolic and Lipidomic Profiles of Tea Plants during Drought Stress and Re-Watering. Front. Plant Sci. 2022, 13, 978531. [Google Scholar] [CrossRef] [PubMed]
  78. Liu, B.; Wang, X.; Li, K.; Cai, Z. Spatially Resolved Metabolomics and Lipidomics Reveal Salinity and Drought-Tolerant Mechanisms of Cottonseeds. J. Agric. Food Chem. 2021, 69, 8028–8037. [Google Scholar] [CrossRef] [PubMed]
  79. Perlikowski, D.; Kierszniowska, S.; Sawikowska, A.; Krajewski, P.; Rapacz, M.; Eckhardt, Ä.; Kosmala, A. Remodeling of Leaf Cellular Glycerolipid Composition under Drought and Re-Hydration Conditions in Grasses from the Lolium-Festuca Complex. Front. Plant Sci. 2016, 7, 1027. [Google Scholar] [CrossRef] [PubMed]
  80. Kränzlein, M.; Schmöckel, S.M.; Geilfus, C.M.; Schulze, W.X.; Altenbuchinger, M.; Hrenn, H.; Roessner, U.; Zörb, C. Lipid Remodeling of Contrasting Maize (Zea mays L.) Hybrids under Repeated Drought. Front. Plant Sci. 2023, 14, 1050079. [Google Scholar] [CrossRef] [PubMed]
  81. Moradi, P.; Mahdavi, A.; Khoshkam, M.; Iriti, M. Lipidomics Unravels the Role of Leaf Lipids in Thyme Plant Response to Drought Stress. Int. J. Mol. Sci. 2017, 18, 2067. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, M.; Shen, Y.; Tao, F.; Yang, S.; Li, W. Submergence Induced Changes of Molecular Species in Membrane Lipids in Arabidopsis thaliana. Plant Divers. 2016, 38, 156–162. [Google Scholar] [CrossRef] [PubMed]
  83. Shao, D.; Yu, C.; Chen, Y.; Qiu, X.; Chen, J.; Zhao, H.; Chen, K.; Wang, X.; Chen, P.; Gao, G.; et al. Lipids Signaling and Unsaturation of Fatty Acids Participate in Ramie Response to Submergence Stress and Hypoxia-Responsive Gene Regulation. Int. J. Biol. Macromol. 2024, 263, 130104. [Google Scholar] [CrossRef] [PubMed]
  84. Xu, Y.; Weng, X.; Tang, S.; Zhang, W.; Zhu, K.; Zhu, G.; Zhang, H.; Wang, Z.; Yang, J. Untargeted Lipidomic Analysis of Milled Rice under Different Alternate Wetting and Soil Drying Irrigation Regimes. J. Integr. Agric. 2024, in press. [CrossRef]
  85. Basu, S.; Kumari, S.; Kumar, P.; Kumar, G.; Rajwanshi, R. Redox Imbalance Impedes Photosynthetic Activity in Rice by Disrupting Cellular Membrane Integrity and Induces Programmed Cell Death under Submergence. Physiol. Plant. 2021, 172, 1764–1778. [Google Scholar] [CrossRef] [PubMed]
  86. Xu, L.; Pan, R.; Zhou, M.; Xu, Y.; Zhang, W. Lipid Remodelling Plays an Important Role in Wheat (Triticum aestivum) Hypoxia Stress. Funct. Plant Biol. 2019, 47, 58–66. [Google Scholar] [CrossRef] [PubMed]
  87. Higashi, Y.; Okazaki, Y.; Takano, K.; Myouga, F.; Shinozaki, K.; Knoch, E.; Fukushima, A.; Saito, K. HEAT INDUCIBLE LIPASE1 Remodels Chloroplastic Monogalactosyldiacylglycerol by Liberating α-Linolenic Acid in Arabidopsis Leaves under Heat Stress. Plant Cell 2018, 30, 1887–1905. [Google Scholar] [CrossRef] [PubMed]
  88. Higashi, Y.; Okazaki, Y.; Myouga, F.; Shinozaki, K.; Saito, K. Landscape of the Lipidome and Transcriptome under Heat Stress in Arabidopsis thaliana. Sci. Rep. 2015, 5, 10533. [Google Scholar] [CrossRef]
  89. Shiva, S.; Samarakoon, T.; Lowe, K.A.; Roach, C.; Vu, H.S.; Colter, M.; Porras, H.; Hwang, C.; Roth, M.R.; Tamura, P.; et al. Leaf Lipid Alterations in Response to Heat Stress of Arabidopsis thaliana. Plants 2020, 9, 845. [Google Scholar] [CrossRef]
  90. Korte, P.; Unzner, A.; Damm, T.; Berger, S.; Krischke, M.; Mueller, M.J. High Triacylglycerol Turnover Is Required for Efficient Opening of Stomata during Heat Stress in Arabidopsis. Plant J. 2023, 115, 81–96. [Google Scholar] [CrossRef]
  91. Qian, W.; Zhu, Y.; Chen, Q.; Wang, S.; Chen, L.; Liu, T.; Tang, H.; Yao, H. Comprehensive Metabolomic and Lipidomic Alterations in Response to Heat Stress during Seed Germination and Seedling Growth of Arabidopsis. Front. Plant Sci. 2023, 14, 1132881. [Google Scholar] [CrossRef] [PubMed]
  92. Sun, A.Z.; Chen, L.S.; Tang, M.; Chen, J.H.; Li, H.; Jin, X.Q.; Yi, Y.; Guo, F.Q. Lipidomic Remodeling in Begonia Grandis Under Heat Stress. Front. Plant Sci. 2022, 13, 843942. [Google Scholar] [CrossRef] [PubMed]
  93. Zhang, Y.; Li, Y.; Han, B.; Liu, A.; Xu, W. Integrated Lipidomic and Transcriptomic Analysis Reveals Triacylglycerol Accumulation in Castor Bean Seedlings under Heat Stress. Ind. Crops Prod. 2022, 180, 114702. [Google Scholar] [CrossRef]
  94. Zoong Lwe, Z.S.; Welti, R.; Anco, D.; Naveed, S.; Rustgi, S.; Narayanan, S. Heat Stress Elicits Remodeling in the Anther Lipidome of Peanut. Sci. Rep. 2020, 10, 22163. [Google Scholar] [CrossRef] [PubMed]
  95. Zhang, X.; Xu, Y.; Huang, B. Lipidomic Reprogramming Associated with Drought Stress Priming-Enhanced Heat Tolerance in Tall Fescue (Festuca arundinacea). Plant Cell Environ. 2019, 42, 947–958. [Google Scholar] [CrossRef]
  96. Hu, H.; Jia, Y.; Hao, Z.; Ma, G.; Xie, Y.; Wang, C.; Ma, D. Lipidomics-Based Insights into the Physiological Mechanism of Wheat in Response to Heat Stress. Plant Physiol. Biochem. 2023, 205, 108190. [Google Scholar] [CrossRef] [PubMed]
  97. Figueira, E.; Matos, D.; Cardoso, P.; Pires, A.; Fernandes, C.; Tauler, R.; Bedia, C. A Biochemical and Lipidomic Approach to Perceive Halimione portulacoides (L.) Response to Mercury: An Environmental Perspective. Mar. Pollut. Bull. 2023, 186, 114393. [Google Scholar] [CrossRef] [PubMed]
  98. Andresen, E.; Lyubenova, L.; Hubáček, T.; Bokhari, S.N.H.; Matoušková, Š.; Mijovilovich, A.; Rohovec, J.; Küpper, H. Chronic Exposure of Soybean Plants to Nanomolar Cadmium Reveals Specific Additional High-Affinity Targets of Cadmium Toxicity. J. Exp. Bot. 2020, 71, 1628–1644. [Google Scholar] [CrossRef] [PubMed]
  99. Wu, S.; Hu, C.; Wang, X.; Wang, Y.; Yu, M.; Xiao, H.; Shabala, S.; Wu, K.; Tan, Q.; Xu, S.; et al. Cadmium-Induced Changes in Composition and Co-Metabolism of Glycerolipids Species in Wheat Root: Glycerolipidomic and Transcriptomic Approach. J. Hazard. Mater. 2022, 423, 127115. [Google Scholar] [CrossRef] [PubMed]
  100. Rastogi, A.; Yadav, D.K.; Szymańska, R.; Kruk, J.; Sedlářová, M.; Pospíšil, P. Singlet Oxygen Scavenging Activity of Tocopherol and Plastochromanol in Arabidopsis thaliana: Relevance to Photooxidative Stress. Plant Cell Environ. 2014, 37, 392–401. [Google Scholar] [CrossRef] [PubMed]
  101. Burgos, A.; Szymanski, J.; Seiwert, B.; Degenkolbe, T.; Hannah, M.A.; Giavalisco, P.; Willmitzer, L. Analysis of Short-Term Changes in the Arabidopsis Thaliana Glycerolipidome in Response to Temperature and Light. Plant J. 2011, 66, 656–668. [Google Scholar] [CrossRef] [PubMed]
  102. Kargiotidou, A.; Deli, D.; Galanopoulou, D.; Tsaftaris, A.; Farmaki, T. Low Temperature and Light Regulate Delta 12 Fatty Acid Desaturases (FAD2) at a Transcriptional Level in Cotton (Gossypium hirsutum). J. Exp. Bot. 2008, 59, 2043–2056. [Google Scholar] [CrossRef] [PubMed]
  103. Callahan, G.; Fillier, T.; Pham, T.H.; Zhu, X.; Thomas, R. The Effects of Clearcut Harvesting on Moss Chloroplast Lipidome and Adaptation to Light Stress during Boreal Forest Regeneration. J. Environ. Manag. 2022, 315, 115126. [Google Scholar] [CrossRef] [PubMed]
  104. Wang, P.; Chen, S.; Gu, M.; Chen, X.; Chen, X.; Yang, J.; Zhao, F.; Ye, N. Exploration of the Effects of Different Blue Led Light Intensities on Flavonoid and Lipid Metabolism in Tea Plants via Transcriptomics and Metabolomics. Int. J. Mol. Sci. 2020, 21, 4606. [Google Scholar] [CrossRef] [PubMed]
  105. Spicher, L.; Almeida, J.; Gutbrod, K.; Pipitone, R.; Dörmann, P.; Glauser, G.; Rossi, M.; Kessler, F. Essential Role for Phytol Kinase and Tocopherol in Tolerance to Combined Light and Temperature Stress in Tomato. J. Exp. Bot. 2017, 68, 5845–5856. [Google Scholar] [CrossRef] [PubMed]
  106. Okazaki, Y.; Otsuki, H.; Narisawa, T.; Kobayashi, M.; Sawai, S.; Kamide, Y.; Kusano, M.; Aoki, T.; Hirai, M.Y.; Saito, K. A New Class of Plant Lipid Is Essential for Protection against Phosphorus Depletion. Nat. Commun. 2013, 4, 1510. [Google Scholar] [CrossRef] [PubMed]
  107. Matich, E.K.; Ghafari, M.; Camgoz, E.; Caliskan, E.; Pfeifer, B.A.; Haznedaroglu, B.Z.; Atilla-Gokcumen, G.E. Time-Series Lipidomic Analysis of the Oleaginous Green Microalga Species Ettlia oleoabundans under Nutrient Stress. Biotechnol. Biofuels 2018, 11, 29. [Google Scholar] [CrossRef] [PubMed]
  108. Custódio, M.; Maciel, E.; Domingues, M.R.; Lillebø, A.I.; Calado, R. Nutrient Availability Affects the Polar Lipidome of Halimione Portulacoides Leaves Cultured in Hydroponics. Sci. Rep. 2020, 10, 6583. [Google Scholar] [CrossRef] [PubMed]
  109. Lowenstein, D.P.; Mayers, K.; Fredricks, H.F.; Van Mooy, B.A.S. Targeted and Untargeted Lipidomic Analysis of Haptophyte Cultures Reveals Novel and Divergent Nutrient-Stress Adaptations. Org. Geochem. 2021, 161, 104315. [Google Scholar] [CrossRef]
  110. Kokabi, K.; Gorelova, O.; Ismagulova, T.; Itkin, M.; Malitsky, S.; Boussiba, S.; Solovchenko, A.; Khozin-Goldberg, I. Metabolomic Foundation for Differential Responses of Lipid Metabolism to Nitrogen and Phosphorus Deprivation in an Arachidonic Acid-Producing Green Microalga. Plant Sci. 2019, 283, 95–115. [Google Scholar] [CrossRef] [PubMed]
  111. Peng, Y.; Lou, H.; Tan, Z.; Ouyang, Z.; Zhang, Y.; Lu, S.; Guo, L.; Yang, B. Lipidomic and Metabolomic Analyses Reveal Changes of Lipid and Metabolite Profiles in Rapeseed during Nitrogen Deficiency. Plant Cell Physiol. 2023, 65, 904–915. [Google Scholar] [CrossRef] [PubMed]
  112. Honda, S.; Yamazaki, Y.; Mukada, T.; Cheng, W.; Chuba, M.; Okazaki, Y.; Saito, K.; Oikawa, A.; Maruyama, H.; Wasaki, J.; et al. Lipidome Profiling of Phosphorus Deficiency-Tolerant Rice Cultivars Reveals Remodeling of Membrane Lipids as a Mechanism of Low P Tolerance. Plants 2023, 12, 1365. [Google Scholar] [CrossRef] [PubMed]
  113. Narasimhan, R.; Wang, G.; Li, M.; Roth, M.; Welti, R.; Wang, X. Differential Changes in Galactolipid and Phospholipid Species in Soybean Leaves and Roots under Nitrogen Deficiency and after Nodulation. Phytochemistry 2013, 96, 81–91. [Google Scholar] [CrossRef] [PubMed]
  114. Liu, M.Y.; Burgos, A.; Ma, L.; Zhang, Q.; Tang, D.; Ruan, J. Lipidomics Analysis Unravels the Effect of Nitrogen Fertilization on Lipid Metabolism in Tea Plant (Camellia sinensis L.). BMC Plant Biol. 2017, 17, 165. [Google Scholar] [CrossRef] [PubMed]
  115. Pfaff, J.; Denton, A.K.; Usadel, B.; Pfaff, C. Phosphate Starvation Causes Different Stress Responses in the Lipid Metabolism of Tomato Leaves and Roots. Biochim. Biophys. Acta Mol. Cell Biol. Lipids 2020, 1865, 158763. [Google Scholar] [CrossRef]
  116. Li, S.; Liu, X.; Yin, L.; Wang, S.; Deng, X. Alteration in Lipid Metabolism Is Involved in Nitrogen Deficiency Response in Wheat Seedlings. Plant Physiol. Biochem. 2024, 214, 108883. [Google Scholar] [CrossRef] [PubMed]
  117. Natera, S.H.A.; Hill, C.B.; Rupasinghe, T.W.T.; Roessner, U. Salt-Stress Induced Alterations in the Root Lipidome of Two Barley Genotypes with Contrasting Responses to Salinity. Funct. Plant Biol. 2016, 43, 207–219. [Google Scholar] [CrossRef] [PubMed]
  118. Ho, W.W.H.; Hill, C.B.; Doblin, M.S.; Shelden, M.C.; van de Meene, A.; Rupasinghe, T.; Bacic, A.; Roessner, U. Integrative Multi-Omics Analyses of Barley Rootzones under Salinity Stress Reveal Two Distinctive Salt Tolerance Mechanisms. Plant Commun. 2020, 1, 100031. [Google Scholar] [CrossRef] [PubMed]
  119. Yu, D.; Boughton, B.A.; Hill, C.B.; Feussner, I.; Roessner, U.; Rupasinghe, T.W.T. Insights Into Oxidized Lipid Modification in Barley Roots as an Adaptation Mechanism to Salinity Stress. Front. Plant Sci. 2020, 11, 1. [Google Scholar] [CrossRef] [PubMed]
  120. Liu, Y.; Wang, L.; Li, X.; Luo, M. Detailed Sphingolipid Profile Responded to Salt Stress in Cotton Root and the GhIPCS1 Is Involved in the Regulation of Plant Salt Tolerance. Plant Sci. 2022, 316, 111174. [Google Scholar] [CrossRef] [PubMed]
  121. Zhang, K.; Lyu, W.; Gao, Y.; Zhang, X.; Sun, Y.; Huang, B. Choline-Mediated Lipid Reprogramming as a Dominant Salt Tolerance Mechanism in Grass Species Lacking Glycine Betaine. Plant Cell Physiol. 2020, 61, 2018–2030. [Google Scholar] [CrossRef] [PubMed]
  122. Guo, Q.; Liu, L.; Rupasinghe, T.W.T.; Roessner, U.; Barkla, B.J. Salt Stress Alters Membrane Lipid Content and Lipid Biosynthesis Pathways in the Plasma Membrane and Tonoplast. Plant Physiol. 2022, 189, 805–826. [Google Scholar] [CrossRef] [PubMed]
  123. Yu, Y.; Kou, M.; Gao, Z.; Liu, Y.; Xuan, Y.; Liu, Y.; Tang, Z.; Cao, Q.; Li, Z.; Sun, J. Involvement of Phosphatidylserine and Triacylglycerol in the Response of Sweet Potato Leaves to Salt Stress. Front. Plant Sci. 2019, 10, 1086. [Google Scholar] [CrossRef] [PubMed]
  124. Pretorius, C.J.; Zeiss, D.R.; Dubery, I.A. The Presence of Oxygenated Lipids in Plant Defense in Response to Biotic Stress: A Metabolomics Appraisal. Plant Signal. Behav. 2021, 16, 1989215. [Google Scholar] [CrossRef] [PubMed]
  125. Bian, Y.; Li, Q.; Zhang, X.; Hao, T.; Liu, N.; Yang, Z.; Yu, J. Lipid Composition Remodeling Plays a Critical Role during the Differential Responses of Leaves and Roots to Heat Stress in Bermudagrass. Environ. Exp. Bot. 2023, 213, 105423. [Google Scholar] [CrossRef]
  126. Han, X.; Yang, Y. Phospholipids in Salt Stress Response. Plants 2021, 10, 2204. [Google Scholar] [CrossRef] [PubMed]
  127. Liang, Y.; Huang, Y.; Liu, C.; Chen, K.; Li, M. Functions and Interaction of Plant Lipid Signalling under Abiotic Stresses. Plant Biol. 2023, 25, 361–378. [Google Scholar] [CrossRef] [PubMed]
  128. Li, H.M.; Yu, C.W. Chloroplast Galactolipids: The Link between Photosynthesis, Chloroplast Shape, Jasmonates, Phosphate Starvation and Freezing Tolerance. Plant Cell Physiol. 2018, 59, 1128–1134. [Google Scholar] [CrossRef] [PubMed]
  129. Song, C.; Yang, Z.; Xiao, X. Membrane Lipid Metabolism Is Involved in Regulating γ-Aminobutyric Acid-Mediated Cold Tolerance in Peach Fruit. Food Front. 2023, 4, 297–307. [Google Scholar] [CrossRef]
  130. Lopes, A.S.; Dias, T.J.; Henschel, J.M.; da Silva, J.H.B.; de Oliveira Sousa, V.F.; Targino, V.A.; da Silva Leal, M.P.; da Silva Gomes, D.; de Albuquerque, M.B.; Batista, D.S. Methyl Jasmonate Mitigates Drought Stress in Purple Basil by Enhancing Photosynthesis and Secondary Metabolism. J. Plant Growth Regul. 2024. [Google Scholar] [CrossRef]
  131. Sato, H.; Mizoi, J.; Shinozaki, K.; Yamaguchi-Shinozaki, K. Complex Plant Responses to Drought and Heat Stress under Climate Change. Plant J. 2024, 117, 1873–1892. [Google Scholar] [CrossRef] [PubMed]
  132. Kim, J.S.; Kidokoro, S.; Yamaguchi-Shinozaki, K.; Shinozaki, K. Regulatory Networks in Plant Responses to Drought and Cold Stress. Plant Physiol. 2024, 195, 170–189. [Google Scholar] [CrossRef] [PubMed]
  133. Bidabadi, S.S.; VanderWeide, J.; Sabbatini, P. Exogenous Melatonin Improves Glutathione Content, Redox State and Increases Essential Oil Production in Two Salvia Species under Drought Stress. Sci. Rep. 2020, 10, 6883. [Google Scholar] [CrossRef] [PubMed]
  134. Kulak, M. Recurrent Drought Stress Effects on Essential Oil Profile of Lamiaceae Plants: An Approach Regarding Stress Memory. Ind. Crops Prod. 2020, 154, 112695. [Google Scholar] [CrossRef]
  135. Sharma, P.; Lakra, N.; Goyal, A.; Ahlawat, Y.K.; Zaid, A.; Siddique, K.H.M. Drought and Heat Stress Mediated Activation of Lipid Signaling in Plants: A Critical Review. Front. Plant Sci. 2023, 14, 1216835. [Google Scholar] [CrossRef] [PubMed]
  136. Cameron, K.D.; Teece, M.A.; Smart, L.B. Increased Accumulation of Cuticular Wax and Expression of Lipid Transfer Protein in Response to Periodic Drying Events in Leaves of Tree Tobacco. Plant Physiol. 2006, 140, 176–183. [Google Scholar] [CrossRef] [PubMed]
  137. Fang, S.; Zhao, P.; Tan, Z.; Peng, Y.; Xu, L.; Jin, Y.; Wei, F.; Guo, L.; Yao, X. Combining Physio-Biochemical Characterization and Transcriptome Analysis Reveal the Responses to Varying Degrees of Drought Stress in Brassica Napus L. Int. J. Mol. Sci. 2022, 23, 8555. [Google Scholar] [CrossRef] [PubMed]
  138. Riyazuddin, R.; Nisha, N.; Singh, K.; Verma, R.; Gupta, R. Involvement of Dehydrin Proteins in Mitigating the Negative Effects of Drought Stress in Plants. Plant Cell Rep. 2022, 41, 519–533. [Google Scholar] [CrossRef] [PubMed]
  139. Sun, M.; Peng, F.; Xiao, Y.; Yu, W.; Zhang, Y.; Gao, H. Exogenous Phosphatidylcholine Treatment Alleviates Drought Stress and Maintains the Integrity of Root Cell Membranes in Peach. Sci. Hortic. 2020, 259, 108821. [Google Scholar] [CrossRef]
  140. Xu, L.; Pan, R.; Zhang, W. Membrane Lipids Are Involved in Plant Response to Oxygen Deprivation. Plant Signal. Behav. 2020, 15, 1771938. [Google Scholar] [CrossRef] [PubMed]
  141. Xie, L.J.; Zhou, Y.; Chen, Q.F.; Xiao, S. New Insights into the Role of Lipids in Plant Hypoxia Responses. Prog. Lipid Res. 2021, 81, 101072. [Google Scholar] [CrossRef] [PubMed]
  142. Bharadwaj, B.; Mishegyan, A.; Nagalingam, S.; Guenther, A.; Joshee, N.; Sherman, S.H.; Basu, C. Physiological and Genetic Responses of Lentil (Lens culinaris) under Flood Stress. Plant Stress 2023, 7, 100130. [Google Scholar] [CrossRef]
  143. Ibrahimova, U.; Kumari, P.; Yadav, S.; Rastogi, A.; Antala, M.; Suleymanova, Z.; Zivcak, M.; Tahjib-Ul-Arif, M.; Hussain, S.; Abdelhamid, M.; et al. Progress in Understanding Salt Stress Response in Plants Using Biotechnological Tools. J. Biotechnol. 2021, 329, 180–191. [Google Scholar] [CrossRef]
  144. Henschel, J.M.; Dias, T.J.; de Moura, V.S.; de Oliveira Silva, A.M.; Lopes, A.S.; da Silva Gomes, D.; Araujo, D.J.; Silva, J.B.M.; da Cruz, O.N.; Batista, D.S. Hydrogen Peroxide and Salt Stress in Radish: Effects on Growth, Physiology, and Root Quality. Physiol. Mol. Biol. Plants 2024. [Google Scholar] [CrossRef]
  145. Lopes, A.S.; Dias, T.J.; Henschel, J.M.; da Silva, T.I.; de Moura, V.S.; Silva, A.M.O.; da Silva Ribeiro, J.E.; Diniz Neto, M.A.; de Oliveira, A.B.; Batista, D.S. Methyl Jasmonate Mitigates Salt Stress and Increases Quality of Purple Basil (Ocimum basilicum L.). S. Afr. J. Bot. 2024, 171, 710–718. [Google Scholar] [CrossRef]
  146. Zhao, S.; Zhang, Q.; Liu, M.; Zhou, H.; Ma, C.; Wang, P. Regulation of Plant Responses to Salt Stress. Int. J. Mol. Sci. 2021, 22, 4609. [Google Scholar] [CrossRef] [PubMed]
  147. Rao, Y.; Peng, T.; Xue, S. Mechanisms of Plant Saline-Alkaline Tolerance. J. Plant Physiol. 2023, 281, 153916. [Google Scholar] [CrossRef] [PubMed]
  148. Fang, S.; Hou, X.; Liang, X. Response Mechanisms of Plants Under Saline-Alkali Stress. Front. Plant Sci. 2021, 12, 667458. [Google Scholar] [CrossRef] [PubMed]
  149. Munns, R.; Passioura, J.B.; Colmer, T.D.; Byrt, C.S. Osmotic Adjustment and Energy Limitations to Plant Growth in Saline Soil. New Phytol. 2020, 225, 1091–1096. [Google Scholar] [CrossRef] [PubMed]
  150. Guo, Q.; Liu, L.; Barkla, B.J. Membrane Lipid Remodeling in Response to Salinity. Int. J. Mol. Sci. 2019, 20, 4264. [Google Scholar] [CrossRef] [PubMed]
  151. Rawat, N.; Singla-Pareek, S.L.; Pareek, A. Membrane Dynamics during Individual and Combined Abiotic Stresses in Plants and Tools to Study the Same. Physiol. Plant. 2021, 171, 653–676. [Google Scholar] [CrossRef] [PubMed]
  152. Li, M.; Yu, A.; Sun, Y.; Hu, Q.; Kang, J.; Chen, L.; Zhu, X.; Yang, Q.; Long, R. Lipid Composition Remodeling and Storage Lipid Conversion Play a Critical Role in Salt Tolerance in Alfalfa (Medicago sativa L.) Leaves. Environ. Exp. Bot. 2023, 205, 105144. [Google Scholar] [CrossRef]
  153. Omoto, E.; Iwasaki, Y.; Miyake, H.; Taniguchi, M. Salinity Induces Membrane Structure and Lipid Changes in Maize Mesophyll and Bundle Sheath Chloroplasts. Physiol. Plant 2016, 157, 13–23. [Google Scholar] [CrossRef] [PubMed]
  154. Helm, L.T.; Shi, H.; Lerdau, M.T.; Yang, X. Solar-Induced Chlorophyll Fluorescence and Short-Term Photosynthetic Response to Drought. Ecol. Appl. 2020, 30, e02101. [Google Scholar] [CrossRef]
  155. Morales-Cedillo, F.; González-Solís, A.; Gutiérrez-Angoa, L.; Cano-Ramírez, D.L.; Gavilanes-Ruiz, M. Plant Lipid Environment and Membrane Enzymes: The Case of the Plasma Membrane H+-ATPase. Plant Cell Rep. 2015, 34, 617–629. [Google Scholar] [CrossRef] [PubMed]
  156. Bejaoui, F.; Salas, J.J.; Nouairi, I.; Smaoui, A.; Abdelly, C.; Martínez-Force, E.; Youssef, N. Ben Changes in Chloroplast Lipid Contents and Chloroplast Ultrastructure in Sulla Carnosa and Sulla Coronaria Leaves under Salt Stress. J. Plant Physiol. 2016, 198, 32–38. [Google Scholar] [CrossRef] [PubMed]
  157. Barkla, B.J.; Garibay-Hernández, A.; Melzer, M.; Rupasinghe, T.W.T.; Roessner, U. Single Cell-Type Analysis of Cellular Lipid Remodelling in Response to Salinity in the Epidermal Bladder Cells of the Model Halophyte Mesembryanthemum crystallinum. Plant Cell Environ. 2018, 41, 2390–2403. [Google Scholar] [CrossRef] [PubMed]
  158. Zuo, Z.F.; Li, Y.; Mi, X.F.; Li, Y.L.; Zhai, C.Y.; Yang, G.F.; Wang, Z.Y.; Zhang, K. Physiological and Lipidomic Response of Exogenous Choline Chloride Alleviating Salt Stress Injury in Kentucky Bluegrass (Poa pratensis). Front. Plant Sci. 2023, 14, 1269286. [Google Scholar] [CrossRef] [PubMed]
  159. Liu, A.; Xiao, Z.; Wang, Z.; Lam, H.M.; Chye, M.L. Galactolipid and Phospholipid Profile and Proteome Alterations in Soybean Leaves at the Onset of Salt Stress. Front. Plant Sci. 2021, 12, 644408. [Google Scholar] [CrossRef]
  160. Yu, W.; Yu, Y.; Wang, C.; Zhang, Z.; Xue, Z. Mechanism by Which Salt Stress Induces Physiological Responses and Regulates Tanshinone Synthesis. Plant Physiol. Biochem. 2021, 164, 10–20. [Google Scholar] [CrossRef] [PubMed]
  161. Sui, N.; Han, G. Salt-Induced Photoinhibition of PSII Is Alleviated in Halophyte Thellungiella Halophila by Increases of Unsaturated Fatty Acids in Membrane Lipids. Acta Physiol. Plant 2014, 36, 983–992. [Google Scholar] [CrossRef]
  162. Salama, K.H.A.; Mansour, M.M.F. Choline Priming-Induced Plasma Membrane Lipid Alterations Contributed to Improved Wheat Salt Tolerance. Acta Physiol. Plant 2015, 37, 170. [Google Scholar] [CrossRef]
  163. Salama, K.H.A.; Mansour, M.M.F.; Ali, F.Z.M.; Abou-Hadid, A.F. NaCl-Induced Changes in Plasma Membrane Lipids and Proteins of Zea mays L. Cultivars Differing in Their Response to Salinity. Acta Physiol. Plant 2007, 29, 351–359. [Google Scholar] [CrossRef]
  164. You, Z.; Zhang, Q.; Peng, Z.; Miao, X. Lipid Droplets Mediate Salt Stress Tolerance in Parachlorella kessleri. Plant Physiol. 2019, 181, 510–526. [Google Scholar] [CrossRef] [PubMed]
  165. Wang, H.; Zhao, S.; Qi, Z.; Yang, C.; Ding, D.; Xiao, B.; Wang, S.; Yang, C. Regulation of Root Exudation in Wheat Plants in Response to Alkali Stress. Plants 2024, 13, 1227. [Google Scholar] [CrossRef] [PubMed]
  166. Wang, H.; Zhao, S.; Sun, B.; Osman, F.M.; Qi, Z.; Ding, D.; Liu, X.; Ding, J.; Zhang, Z. Carboxylic Acid Accumulation and Secretion Contribute to the Alkali-Stress Tolerance of Halophyte Leymus Chinensis. Front. Plant Sci. 2024, 15, 1366108. [Google Scholar] [CrossRef] [PubMed]
  167. Skorzynska-Polit, E. Lipid Peroxidation in Plant Cells, Its Physiological Role and Changes under Heavy Metal Stress. Acta Soc. Bot. Pol. 2007, 76, 49–54. [Google Scholar] [CrossRef]
  168. Zhang, F.Q.; Wang, Y.S.; Lou, Z.P.; Dong, J. De Effect of Heavy Metal Stress on Antioxidative Enzymes and Lipid Peroxidation in Leaves and Roots of Two Mangrove Plant Seedlings (Kandelia candel and Bruguiera gymnorrhiza). Chemosphere 2007, 67, 44–50. [Google Scholar] [CrossRef] [PubMed]
  169. Devi, S.R.; Prasad, M.N.V. Membrane Lipid Alterations in Heavy Metal Exposed Plants. In Heavy Metal Stress in Plants; Springer: Berlin/Heidelberg, Germany, 1999. [Google Scholar] [CrossRef]
  170. de Bang, T.C.; Husted, S.; Laursen, K.H.; Persson, D.P.; Schjoerring, J.K. The Molecular–Physiological Functions of Mineral Macronutrients and Their Consequences for Deficiency Symptoms in Plants. New Phytol. 2021, 229, 2446–2469. [Google Scholar] [CrossRef] [PubMed]
  171. Francis, B.; Aravindakumar, C.T.; Brewer, P.B.; Simon, S. Plant Nutrient Stress Adaptation: A Prospect for Fertilizer Limited Agriculture. Environ. Exp. Bot. 2023, 213, 105431. [Google Scholar] [CrossRef]
  172. Terrer, C.; Jackson, R.B.; Prentice, I.C.; Keenan, T.F.; Kaiser, C.; Vicca, S.; Fisher, J.B.; Reich, P.B.; Stocker, B.D.; Hungate, B.A.; et al. Nitrogen and Phosphorus Constrain the CO2 Fertilization of Global Plant Biomass. Nat. Clim. Chang. 2019, 9, 684–689. [Google Scholar] [CrossRef]
  173. Tränkner, M.; Tavakol, E.; Jákli, B. Functioning of Potassium and Magnesium in Photosynthesis, Photosynthate Translocation and Photoprotection. Physiol. Plant 2018, 163, 414–431. [Google Scholar] [CrossRef] [PubMed]
  174. Lan, P.; Li, W.; Schmidt, W. “Omics” Approaches Towards Understanding Plant Phosphorus Acquisition and Use. In Phosphorus Metabolism in Plants; Wiley Blackwell: New York, NY, USA, 2015; Volume 48, pp. 65–98. ISBN 9781118958841. [Google Scholar]
  175. Singer, S.D.; Zou, J.; Weselake, R.J. Abiotic Factors Influence Plant Storage Lipid Accumulation and Composition. Plant Sci. 2016, 243, 1–9. [Google Scholar] [CrossRef] [PubMed]
  176. Yavari, N.; Tripathi, R.; Wu, B.-S.; MacPherson, S.; Singh, J.; Lefsrud, M. The Effect of Light Quality on Plant Physiology, Photosynthetic, and Stress Response in Arabidopsis thaliana Leaves. PLoS ONE 2021, 16, e0247380. [Google Scholar] [CrossRef] [PubMed]
  177. Lakshmanan, M.; Lim, S.H.; Mohanty, B.; Kim, J.K.; Ha, S.H.; Lee, D.Y. Unraveling the Light-Specific Metabolic and Regulatory Signatures of Rice through Combined in Silico Modeling and Multiomics Analysis. Plant Physiol. 2015, 169, 3002–3020. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Publication metrics regarding plant lipidomics. Research was performed as follows: term “Plant AND lipidomics”, limiting search to articles in English language (a,c,e), and term “Plant AND stress AND lipidomics”, limiting search to articles in English language (b,d,f). Total findings were divided by year of publication (a,b), subject area (c,d), or country (e,f).
Figure 1. Publication metrics regarding plant lipidomics. Research was performed as follows: term “Plant AND lipidomics”, limiting search to articles in English language (a,c,e), and term “Plant AND stress AND lipidomics”, limiting search to articles in English language (b,d,f). Total findings were divided by year of publication (a,b), subject area (c,d), or country (e,f).
Agronomy 14 01670 g001
Table 1. An overview of the main lipidomic studies approaching plant abiotic stress responses.
Table 1. An overview of the main lipidomic studies approaching plant abiotic stress responses.
Stress ConditionPlant SpeciesTissueAnalytical TechniqueReference
AlkaliMaizeRootsESI-MS/MSXu et al. [55]
SunflowerSeedsLC-MSLu et al. [56]
ColdArabidopsis thalianaLeavesLC-MSTarazona et al. [57]
Arabidopsis thalianaLeavesGC-MSChen et al. [58]
BananaFruitsLC-MSLiu et al. [59]
BarleyLeavesLC-MSZhao et al. [60]
Bell pepperFruitsESI-MS/MSKong et al. [61]
Bell pepperFruitsLC-MSXu et al. [62]
BlueberryFruitsGC-MSWang et al. [63]
CabbageLeavesLC-MSZainal et al. [64]
CottonSeedsLC-MSDhaliwal et al. [65]
DendrobiumLeavesUPLC-QTOF-MSZhan et al. [66]
MaizeRootsESI-MS/MSZhao et al. [67]
MaizeLeavesLC-MS/MSGao et al. [68]
Meconopsis racemosaLeavesESI-MS/MSZheng et al. [69]
Moss speciesLeavesUPLC-QTOF-MSLu et al. [70]
Panax notoginsengLeavesESI-MS/MSLiu et al. [71]
PeachFruitsGC-MSZhu et al. [72]
Peach FruitsESI-MS/MSChen et al. [14]
Wheat LeavesLC-QqQ-MSCheong et al. [73]
Wild tomatoLeavesUPLC-QTOF-MSZhang et al. [74]
DroughtArabidopsis thalianaLeavesLC-MS/MSTarazona et al. [57]
Argan treeLeavesGC-MSRabeh et al. [75]
BarleyRootsLC-MSSarabia et al. [76]
Camellia sinensisLeavesGC-MS and LC-MSShen et al. [77]
CottonSeedsLC-MS/MSLiu et al. [78]
Italian ryegrass × tall fescueLeavesUPLC-MSPerlikowski et al. [79]
MaizeLeavesLC-MSKränzlein et al. [80]
SorghumLeaf cuticleGC-MSZhang et al. [49]
ThymeLeavesDIFT-ICR-MSMoradi et al. [81]
WaterloggingArabidopsis thalianaLeavesESI-MS/MSWang et al. [82]
Boehmeria niveaLeavesUPLC-QTOF/MSShao et al. [83]
RiceGrainsLC-MSXu et al. [84]
RiceLeavesLC-MSBasu et al. [85]
WheatLeavesESI-MS/MSXu et al. [86]
HeatArabidopsis thalianaLeavesLC-MSHigashi et al. [87]
Arabidopsis thalianaLeavesLC-MSHigashi et al. [88]
Arabidopsis thalianaLeavesLC-MSShiva et al. [89]
Arabidopsis thalianaLeavesLC-MSKorte et al. [90]
Arabidopsis thalianaSeeds and seedlingsESI/MSQian et al. [91]
BegoniaLeaves LC-MSSun et al. [92]
Castor beanLeavesUPLC-QTOF/MSZhang et al. [93]
Panax notoginsengLeavesESI-MS/MSLiu et al. [71]
PeanutAnthersESI-MS/MSLwe et al. [94]
SoybeanLeavesESI-MS/MSNarayanan et al. [48]
Tall fescueLeavesESI-MS/MSZhang et al. [95]
Tall fescueLeavesESI-MS/MSZhang et al. [54]
WheatLeavesLC-MS/MSHu et al. [96]
Heavy metalsHalimione portulacoidesLeaves, stems, and rootsLC-MSFigueira et al. [97]
Soybean LeavesLC-MSAndresen et al. [98]
Wheat RootsESI-MS/MSWu et al. [99]
LightArabidopsis thalianaLeavesHPLCRastogi et al. [100]
Arabidopsis thalianaLeavesUPLC-MSBurgos et al. [101]
CottonLeaves and rootsGC-MSKargiotidou et al. [102]
Moss LeavesLC-MSCallahan et al. [103]
Tea plantsLeavesUPLC-TOF-MSWang et al. [104]
TomatoLeavesUHPLC-APCI-QTOF-MSSpicher et al. [105]
Nutrient deficiencyArabidopsis thalianaLeavesLC-MSOkazaki et al. [106]
Ettlia oleoabundansCellsLC–QTOF-MSMatich et al. [107]
Halimione portulacoidesLeavesLC-MSCustódio et al. [108]
HaptophytesCellsHPLC–HRAM–MSLowenstein et al. [109]
Lobosphaera incisaCellsLC-MSKokabi et al. [110]
RapeseedLeavesLC-MS/MSPeng et al., 2023 [111]
RiceLeaves and rootsUPLC-QTOF-MSHonda et al. [112]
SoybeanLeaves and rootsESI-MS/MSNarasimhan et al. [113]
Tea plantLeavesUPLC-MSLiu et al. [114]
TomatoLeaves and rootsGC-MSPfaff et al. [115]
WheatLeavesGC-MSLi et al. [116]
SaltBarleyRootsHPLC-ESI-QTOF-MSNatera et al. [117]
BarleyRootsHPLC-ESI-QTOF-MSHo et al. [118]
BarleyRootsHPLC-ESI-QTOF-MSYu et al. [119]
Carex rigescensLeavesESI-MS/MSHu et al. [51]
CottonSeedsLC-MS/MS Liu et al. [78]
Cotton RootsLC-MSLiu et al. [120]
Kentucky bluegrassLeavesESI-MS/MSZhang et al. [121]
Mesembryanthemum crystallinumLeavesHPLC–ESI–MS/MSGuo et al. [122]
Sweet PotatoLeavesUPLC-QTOF/MSYu et al. [123]
Tomato LeavesUHPLC/QTOF-MSBuffagni et al. [52]
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

Henschel, J.M.; Andrade, A.N.d.; dos Santos, J.B.L.; da Silva, R.R.; da Mata, D.A.; Souza, T.; Batista, D.S. Lipidomics in Plants Under Abiotic Stress Conditions: An Overview. Agronomy 2024, 14, 1670. https://doi.org/10.3390/agronomy14081670

AMA Style

Henschel JM, Andrade ANd, dos Santos JBL, da Silva RR, da Mata DA, Souza T, Batista DS. Lipidomics in Plants Under Abiotic Stress Conditions: An Overview. Agronomy. 2024; 14(8):1670. https://doi.org/10.3390/agronomy14081670

Chicago/Turabian Style

Henschel, Juliane Maciel, Antônio Nunes de Andrade, Josefa Bruna Lima dos Santos, Rodrigo Ribeiro da Silva, Djair Alves da Mata, Tancredo Souza, and Diego Silva Batista. 2024. "Lipidomics in Plants Under Abiotic Stress Conditions: An Overview" Agronomy 14, no. 8: 1670. https://doi.org/10.3390/agronomy14081670

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop