Next Article in Journal
Integrative Approaches to Myopathies and Muscular Dystrophies: Molecular Mechanisms, Diagnostics, and Future Therapies
Previous Article in Journal
Multiplexed Quantification of First-Trimester Serum Biomarkers in Healthy Pregnancy
Previous Article in Special Issue
Phylogenetic Determinants Behind the Ecological Traits of Relic Tree Family Juglandaceae, Their Root-Associated Symbionts, and Response to Climate Change
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in the Biosynthetic Regulation and Functional Mechanisms of Glycine Betaine for Enhancing Plant Stress Resilience

1
National Key Laboratory for the Development and Utilization of Forest Food Resources, Co-Innovation Center for Sustainable Forestry in Southern China, State Key Laboratory of Tree Genetics and Breeding, Key Laboratory of State Forestry and Grassland Administration on Subtropical Forest Biodiversity Conservation, College of Life Sciences, Nanjing Forestry University, Nanjing 210037, China
2
Laboratory Center of Life Sciences, College of Life Sciences, Nanjing Agricultural University, Nanjing 210095, China
*
Authors to whom correspondence should be addressed.
These authors contribute equally to this work.
Int. J. Mol. Sci. 2025, 26(16), 7971; https://doi.org/10.3390/ijms26167971
Submission received: 23 June 2025 / Revised: 8 August 2025 / Accepted: 13 August 2025 / Published: 18 August 2025
(This article belongs to the Collection Advances in Molecular Plant Sciences)

Abstract

Plants are frequently exposed to a range of abiotic stresses, including drought, salinity, extreme temperatures, and heavy metals, that severely impair their growth and productivity. Among the adaptive mechanisms that plants have evolved, the accumulation of glycine betaine (GB), a naturally occurring, zwitterionic, and chemically stable osmoprotectant, has been widely recognized as a key strategy for stress tolerance. In higher plants, GB is primarily synthesized via the two-step oxidation of choline, catalyzed by choline monooxygenase (CMO) and betaine aldehyde dehydrogenase (BADH). GB contributes to cellular homeostasis by modulating osmotic balance, regulating ion flux, scavenging reactive oxygen species (ROS), enhancing antioxidant defense systems, and stabilizing proteins and membrane structures. Both exogenous application of GB and genetic engineering approaches aimed at enhancing endogenous GB biosynthesis have been shown to significantly improve plant tolerance to a variety of abiotic stresses. In this review, we provide a comprehensive overview of recent advances in the understanding of GB biosynthesis, its regulatory mechanisms, and its multifaceted roles in plant stress responses. We also highlight emerging prospects for leveraging GB-centered strategies to enhance crop resilience in challenging environmental conditions.

1. Introduction

Global climate change has markedly increased the frequency and intensity of abiotic stresses, including drought, salinity, and extreme temperatures, which impose continuous challenges on plant systems, severely limiting their growth, development, and overall productivity [1]. These environmental stressors contribute substantially to global agricultural yield losses. In response, plants have evolved a range of adaptive mechanisms, among which the accumulation of compatible solutes, such as amino acids and their derivatives, plays a pivotal role in protecting cellular function under adverse conditions [2].
Glycine betaine (GB), chemically defined as N,N,N-trimethylglycine, is a classic compatible solute [3]. Its zwitterionic structure, comprising a positively charged trimethylammonium group and a negatively charged carboxyl group, confers high solubility and chemical stability. GB can accumulate at substantial concentrations within plant cells without interfering with normal metabolic processes, while significantly enhancing tolerance to osmotic stress, high- and low-temperature stress, and oxidative damage [4,5]. In recent years, GB has attracted increasing attention as a multifunctional protective agent in plant stress physiology. A comprehensive understanding of its biosynthetic pathways, regulatory networks, and physiological roles is essential for developing sustainable strategies to improve crop resilience. Such knowledge could inform the targeted use of GB in agriculture and environmental management, advancing efforts to mitigate the impact of abiotic stresses in a changing climate.

2. Biosynthetic Pathways and Regulatory Mechanisms of GB in Plants

2.1. Choline Oxidation Pathway

The choline oxidation pathway constitutes the primary route for GB biosynthesis in most higher plants (Figure 1A). This pathway proceeds via two sequential enzymatic steps. First, choline is hydroxylated to betaine aldehyde by choline monooxygenase (CMO), a stress-inducible, ferredoxin-dependent monooxygenase localized in the chloroplast stroma [6,7]. For example, in salt- and drought-tolerant species like sugar beet (Beta vulgaris) and spinach, CMO transcript and protein levels increase by 3- to 5-fold during drought and up to 7-fold under high salinity (400 mM NaCl), highlighting its pivotal role in initiating GB biosynthesis in response to abiotic stress [8].
CMO requires Fe2+, O2, and NADPH as cofactors (Figure 1A). In photosynthetic electron transport, ferredoxin (Fd) receives electrons from PSI and donates them to ferredoxin-NADP+ reductase (FNR), which reduces NADP+ to NADPH, an essential electron donor for CMO activity [9,10]. As a result, CMO function is tightly coupled to the efficiency of photosynthesis. Indeed, stress-induced damage to the photosynthetic apparatus can impair ferredoxin reduction and NADPH production, thereby hindering GB synthesis [11]. This connection highlights that effective GB biosynthesis requires an intact photosynthetic system.
Following betaine aldehyde formation, the second step of GB biosynthesis involves its irreversible oxidation to GB, catalyzed by betaine aldehyde dehydrogenase (BADH), an NAD+/NADP+-dependent enzyme predominantly localized in the chloroplast stroma [12,13]. Notably, plant BADHs generally exhibit higher catalytic activity with NAD+ as cofactor [12,14,15], allowing partial functional independence from light-driven reducing power. Overexpression of BADH strongly enhances GB accumulation and confers improved abiotic stress tolerance. For instance, transgenic wheat lines overexpressing HvBADH1 accumulated up to 21.82-fold more GB under salinity stress (150 mM NaCl) [16], while OsBADH1-overexpressing rice lines similarly exhibited increased GB content and enhanced salt tolerance [17]. Conversely, BADH-deficient mutants show severely impaired GB biosynthesis and heightened stress sensitivity. These findings underscore BADH as a critical determinant of GB-mediated stress tolerance in plants.
In most plants, GB biosynthesis proceeds through the choline oxidation pathway. However, in some microorganisms, alternative enzymes mediate the initial oxidation of choline. For example, in certain Gram-negative bacteria (e.g., Acinetobacter baumannii), choline is oxidized by choline dehydrogenase (CDH), known as BetA [18]. Other microbes, such as the fungus Arthrobacter globiformis, employ choline oxidase (COX) for the same purpose [19]. Nevertheless, BADH remains the conserved enzyme responsible for the subsequent conversion of betaine aldehyde to GB in both plant and microbial systems.
Importantly, not all plants possess a fully functional GB biosynthetic pathway (Figure 1B). Genomic analyses of Oryza sativa (rice) and Arabidopsis thaliana reveal divergence in their CMO and BADH gene families. Rice contains a CMO and at least two BADH homologs. The rice CMO localizes to the chloroplast and is likely responsible for the initial oxidation of choline [20]. Among its BADH homologs, OsBADH1 is peroxisomal and may oxidize acetaldehyde [21], while OsBADH2 is implicated in aroma biosynthesis [22]. In Arabidopsis, the CMO homolog AtCMO is weakly expressed, and two BADH-like genes, AtALDH10A8 and AtALDH10A9, encode enzymes localized to chloroplasts and peroxisomes, respectively, both capable of oxidizing betaine aldehyde [23,24,25]. Crucially, both Arabidopsis and rice accumulate negligible levels of GB under normal or stress conditions due to low CMO protein abundance [24]. Overexpression studies have shown that introducing CMO alone is insufficient for substantial GB production unless sufficient BADH activity is also present [26]. For example, tobacco engineered with expressing SpCMO (from spinach) failed to accumulate GB unless co-expressed with a functional BADH gene [27]. Therefore, successful metabolic engineering of GB biosynthesis in non-accumulator plants requires coordinated expression of both CMO and BADH [24]. Indeed, transgenic Arabidopsis [28], rapeseed [24], tobacco (Nicotiana tabacum) [29], and potato [30] expressing CMO and BADH together exhibited significantly elevated GB levels and improved stress tolerance.
The transcription of CMO and BADH is regulated by a complex array of cis-regulatory elements (CREs) responsive to environmental and hormonal signals [31]. Bioinformatic analyses of promoter regions in rice and Arabidopsis have identified enrichment of abscisic acid-responsive elements (ABREs), dehydration-responsive elements (DREs), low-temperature responsive elements (LTREs), and antioxidant response elements (AREs), suggesting that both genes are tightly integrated into the convergence of stress-signaling networks [32,33]. Additionally, promoters often contain light-responsive motifs (e.g., G-box, GT1-motif), tissue-specific elements (e.g., ROOTMOTIF), and hormone-associated sites (e.g., TGA-element, CGTCA-motif), reflecting intricate spatial and temporal regulation.
DNA methylation and histone acetylation are key epigenetic mechanisms that regulate chromatin architecture, mediating long-lasting changes in gene accessibility and transcriptional activity [34,35,36,37]. Increasing evidence shows that these modifications, including DNA methylation and diverse histone marks, critically regulate stress-responsive gene expression in plants [38]. Although direct evidence linking GB biosynthetic genes, such as BADH and CMO, and epigenetic regulation remains elusive, this connection is a plausible and compelling hypothesis, especially under abiotic stress conditions where rapid, reversible gene regulation is vital.
Beyond gene regulation, substrate availability, particularly of choline, is a major control point in GB biosynthesis. In transgenic tobacco overexpressing SpCMO, GB accumulation was limited by insufficient endogenous choline levels. Supplementation with 5 mM choline or phosphocholine significantly enhanced both intracellular choline and GB content [27]. Choline supply in plants originates from either de novo synthesis or phospholipid degradation. De novo biosynthesis involves sequential methylation of phosphoethanolamine by phosphoethanolamine N-methyltransferase (PEAMT), producing phosphocholine (PCho), which is then dephosphorylated to yield free choline [39]. In Beta vulgaris, phospholipid turnover and the cytidine diphosphate (CDP)-choline pathway constitute the major sources of choline. This pathway involves PEAMT-mediated PCho formation, followed by conversion to CDP-choline via phosphocholine cytidylyltransferase (CCT), and its incorporation into phosphatidylcholine by choline phosphotransferase (CPT). Subsequent hydrolysis of phosphatidylcholine by phospholipases releases free choline for GB synthesis [40]. In contrast, non-accumulator plants like tobacco rely primarily on the PEAMT pathway. Overexpression of PEAMT in tobacco increased phosphocholine levels 5-fold and free choline by 50-fold without affecting membrane lipid content or plant growth [39], underscoring the feasibility of enhancing substrate flux to support engineered GB biosynthesis.
Although GB biosynthesis, particularly via the CMO pathway, requires reducing power derived from photosystem I (PSI), GB itself plays a protective role in maintaining photosynthetic efficiency, especially the function of photosystem II (PSII), under stress conditions. This apparent paradox is addressed by several regulatory and physiological features that allow GB synthesis to be sustained even when photosynthesis is partially impaired.
Under abiotic stress (e.g., salinity, drought, cold), PSII is more vulnerable than PSI, which enables limited electron flow through PSI to persist and maintain CMO activity. Moreover, GB specifically stabilizes the oxygen-evolving complex of PSII, protects D1 protein from degradation, and promotes its turnover, thereby facilitating PSII repair and recovery [41,42]. This creates a positive feedback loop, in which GB biosynthesis initially relies on PSI-derived electrons, but the accumulated GB helps restore or preserve PSII activity, indirectly supporting sustained photosynthetic function and further GB production.
In addition, the second step of GB synthesis, catalyzed by BADH, can utilize NAD+, a cofactor not directly linked to photosynthetic electron flow, indicating a partial uncoupling from PSI under stress. Experimental evidence further supports this flexibility: in sugar beet and sorghum, the expression of CMO and BADH is strongly upregulated under salinity and drought, leading to continued GB accumulation despite reduced carbon fixation and water availability [8,43]. In Suaeda liaotungensis, stress-inducible elements in the BADH promoter region further enhance its transcription in response to NaCl [44], suggesting that transcriptional regulation compensates for the transient suppression of CMO activity during acute stress. These findings together suggest that while GB synthesis depends on photosynthesis for reducing power, the pathway exhibits regulatory plasticity and resilience, allowing for continued GB accumulation that in turn supports and protects the photosynthetic machinery, a finely tuned feedback mechanism crucial for stress adaptation.

2.2. Glycine Methylation Pathway

An alternative GB biosynthetic pathway exists in certain halophilic microorganisms such as Ectothiorhodospira halochloris and Actinopolyspora halophila, where GB is synthesized via the glycine methylation pathway rather than choline oxidation. In this pathway, glycine undergoes three successive S-adenosylmethionine (AdoMet)-dependent methylation reactions, yielding sarcosine, dimethylglycine (DMG), and ultimately GB [45] (Figure 1B). This multi-step pathway was successfully transferred to Escherichia coli: heterologous expression of microbial genes encoding glycine betaine methyltransferase (GSMT) and sarcosine/dimethylglycine methyltransferase (SDMT) enabled GB synthesis in E. coli even without choline, greatly improving salt tolerance [46,47]. These findings underscore the biotechnological potential of the glycine methylation pathway as an alternative route for GB biosynthesis. Its independence from choline and compatibility with microbial chassis systems make it a promising candidate for synthetic biology applications aimed at improving abiotic stress tolerance in plants and microbes.
No native homologs of these methyltransferases are known in higher plants. However, expressing microbial GSMT and SDMT in plants that cannot synthesize GB represents a novel strategy for enabling them to produce GB. For example, Waditee et al. (2005) engineered Arabidopsis by introducing ApGSMT and ApSDMT from the halophilic archaeon Aphanothece halophytica [48]. The transgenic plants accumulated significant GB in roots, stems, leaves, and flowers, levels exceeding those in lines expressing only choline-oxidizing enzymes. Likewise, Ref. [49] introduced Mpgsmt and Mpsdmt from Methanohalophilus portucalensis into Arabidopsis, successfully enabling betaine production in plants. These examples demonstrate that the glycine methylation pathway can be harnessed to confer GB biosynthesis in crops lacking the native choline oxidation pathway.
Beyond plants, mammals and certain microorganisms also possess choline oxidation systems functionally analogous to those found in plants. In some Gram-negative bacteria, such as Acinetobacter baumannii, choline oxidation is catalyzed by choline dehydrogenase (BetA) [18]. In E. coli, choline is converted to GB through the sequential actions of choline dehydrogenase (CDH), encoded by betA, and BADH, encoded by betB [50,51]. In contrast, certain eukaryotic microorganisms and fungi, such as Arthrobacter globiformis, employ a distinct enzyme, choline oxidase (COX), to catalyze this process [19].
Additionally, GB may also be synthesized through stress-induced pathways (Figure 1C), potentially originating from serine derived via (i) the non-phosphorylated glycerate pathway, (ii) the phosphorylated phosphohydroxypyruvate pathway [52,53], or (iii) the salt stress-induced photorespiratory glycerate pathway [54,55], as proposed by Annunziata [56]. Serine is subsequently converted into Ethanolamine via serine decarboxylase. These pathways link primary metabolism with osmoprotectant biosynthesis under abiotic stress.

2.3. Mechanisms of GB Uptake and Transporters in Plants

Efficient GB biosynthesis, whether from endogenous metabolism or exogenous application, depends not only on the availability of choline but also on its intracellular transport, particularly into the chloroplast where GB synthesis occurs. Cytosolic choline, whether synthesized de novo or absorbed from the environment, must be effectively delivered to the sites of GB biosynthesis.
Plants possess multiple transport systems for choline. These include a high-affinity choline transporter (CHT1), choline transporter-like (CTL) proteins of intermediate affinity, and low-affinity organic cation transporters (OCTs) [57]. The first plant choline transporter identified was AtCTL1 from Arabidopsis thaliana, which mediates choline uptake. Intriguingly, AtCTL1 also influences auxin distribution by promoting the trafficking of auxin efflux carriers, linking choline transport to developmental processes. The Arabidopsis CTL1 ortholog in roots, known as Choline Transporter1 (CHER1), may similarly facilitate choline transport to maintain root growth [58]. Therefore, GB production depends not only on the amount of cytosolic choline but also on how efficiently it is imported into chloroplasts [59,60]. Supporting this, Yamada et al. (2011) used a yeast mutant deficient in choline uptake to show that sugar beet Proline/GB transporters (Pro/BeTs) have a higher affinity for choline than for GB [61]. These findings highlight a crucial layer of regulation in GB metabolism at the level of membrane transport, linking substrate availability with organellar import and stress-induced demand.
In bacteria, the osmoprotective choline transporter BetT provides a model. E. coli uses BetT for choline uptake, enabling subsequent GB synthesis. Yang et al. (2024) solved the cryo-EM structure of BetT, revealing how it binds choline [62]. Such structural insights could guide engineering of plant transport systems to improve GB accumulation. To date, no plant protein has been identified that specifically transports GB across membranes, and the H+-coupled transporters that might move GB remain to be discovered [42,63].
At the whole-plant level, experiments show that GB applied to one part can move to others. Park et al. (2006) demonstrated that applying GB to a mature tomato leaf resulted in rapid transport of GB via the phloem to meristematic tissues, such as flower buds and shoot apices [64]. Similarly, GB-accumulating transgenic Arabidopsis [65] and tomato [66] plants actively translocated foliar GB through the phloem to growing sinks [64,66]. This phloem mobility suggests that exogenous GB can distribute throughout the plant, reaching target tissues to confer stress protection where it is most needed.

3. Functional Roles of GB in Plant Stress Tolerance Mechanisms

GB enhances plant tolerance to diverse abiotic stresses through a suite of coordinated physiological and molecular mechanisms (Figure 2). Under stress conditions, GB modulates key phytohormonal signaling pathways, including abscisic acid (ABA), gibberellins (GA), and salicylic acid (SA), thereby regulating stress-responsive gene expression and adaptive responses. It contributes to osmotic and ionic homeostasis by facilitating Na+ exclusion and promoting the accumulation of compatible solutes such as proline and soluble sugars. GB also strengthens antioxidant defenses by activating enzymes like superoxide dismutase (SOD) and catalase (CAT) and by enhancing the ascorbate-glutathione (AsA-GSH) cycle for efficient reactive oxygen species (ROS) detoxification. Furthermore, GB helps preserve photosynthetic efficiency and membrane stability through transcriptional regulation and hormone-mediated crosstalk. Notably, GB accumulation may also deter herbivory by interfering with insect feeding behavior, suggesting a broader role in enhancing plant resilience under combined abiotic and biotic stresses (Figure 2).

3.1. Maintenance of Cellular Homeostasis

3.1.1. Regulation of Osmotic Adjustment

Osmotic adjustment is one of the earliest and most critical cellular responses to abiotic stress [67]. Under conditions such as salinity and drought, the external water potential declines, leading to cellular dehydration and turgor loss. GB, due to its high solubility and metabolic inertness, accumulates in the cytosol without interfering with normal biochemical processes [68,69]. Its accumulation enables cells to retain water and maintain turgor pressure, thereby sustaining metabolic activity and physiological function under osmotic stress.
Moreover, GB can stimulate the accumulation of other compatible solutes such as proline and soluble sugars. Under salt stress, exogenous GB has been shown to upregulate the expression of pyrroline-5-carboxylate synthetase (P5CS), a key gene in proline biosynthesis, and enhance sucrose phosphate synthase (SPS) activity, resulting in elevated levels of proline and sugars [70]. The combined effect of GB, proline, and sugars synergistically reduces cellular osmotic potential and helps maintain turgor under drought and salinity stress. Thus, GB functions not only as a primary osmolyte but also as a regulatory molecule that coordinates the synthesis of additional osmoprotectants to support water balance in stressed plants.

3.1.2. Regulation of Ion Homeostasis

Maintaining ion homeostasis is essential for cellular function, particularly under salt stress, where excessive accumulation of Na+ and Cl can impair enzymatic activity and disrupt membrane potentials. GB contributes to ionic balance by enhancing the selective uptake of beneficial ions such as K+ and promoting the exclusion or vacuolar compartmentalization of toxic ions [71]. This regulatory effect helps preserve cellular ionic equilibrium and supports metabolic stability under saline conditions.
At the molecular level, GB treatment upregulates the expression of potassium transporter genes. For instance, members of the K+ uptake permease (KUP) family (e.g., Potassium transporter 21-like) are induced by GB, enhancing both high- and low-affinity K+ uptake [72]. This helps maintain an optimal cytosolic K+ concentration and a favorable K+/Na+ ratio, which is essential for enzyme function and cell turgor. Under salt stress, GB-treated plants tend to accumulate more K+ in roots and restrict Na+ transport to shoots [73]. Additionally, GB can activate H+/K+ co-transport systems and stimulate Na+ efflux pumps, further reducing Na+ toxicity. Through these actions, GB helps maintain stable ion gradients and preserves metabolic and signaling functions during salt stress.

3.1.3. Regulation of Redox Homeostasis

Redox homeostasis entails a dynamic balance between ROS production and antioxidant defense mechanisms [74]. Abiotic stresses frequently disrupt this balance, leading to excessive ROS accumulation and oxidative damage to lipids, proteins, and DNA. GB contributes to redox regulation both by directly scavenging ROS and by enhancing endogenous antioxidant systems. For example, exogenous GB application in rice seedlings significantly increased glutathione reductase (GR) and ascorbate peroxidase (APX) activities by 30% and 25%, respectively, thereby reinforcing the ascorbate-glutathione (AsA-GSH) cycle and limiting H2O2 accumulation [75]. In pepper, GB treatment upregulated genes encoding superoxide dismutase (CaSOD), peroxidase (CaPOD), catalase (CaCAT), CaGR1, and dehydroascorbate reductase (CaDHAR) [76]. Through these mechanisms, GB mitigates oxidative stress and helps maintain redox homeostasis under adverse environmental conditions.

3.1.4. Stabilization of Membrane Structures

Membrane stability is essential for maintaining cellular compartmentalization, transport, and signal transduction [77]. Under abiotic stress, lipid peroxidation and phase transitions can disrupt membrane architecture and impair functionality. GB helps preserve membrane stability by interacting with lipid bilayers, enhancing their fluidity and thermal resilience. In chloroplasts, GB maintains the fluidity of thylakoid membranes and prevents deleterious phase changes under heat or salinity stress. This stabilization effect protects PSII complexes, thereby supporting sustained electron transport and maintaining photosynthetic efficiency under adverse conditions [78].
GB similarly stabilizes mitochondrial membranes: it maintains crista architecture and prevents the collapse of membrane potential and disassembly of respiratory complexes [75,79]. These actions ensure efficient ATP synthesis under stress. At the molecular level, GB is thought to stabilize membrane-associated proteins as well, thereby preserving ion gradients, enzyme activities, and signaling processes essential for survival under adverse conditions. By reinforcing membrane stability across organelles, GB helps safeguard cellular structures during stress.

3.2. Regulation of Metabolic Networks

Beyond maintaining cellular homeostasis, GB modulates broader metabolic networks to support plant performance under stress. Under high temperatures or salt stress, GB mitigates photoinhibition and enhances photosynthetic efficiency by stabilizing photosynthetic enzymes and membrane systems. Specifically, GB specifically interacts with thylakoid lipids to stabilize PSII reaction centers, thereby preserving the structural integrity and functionality of the photosynthetic apparatus [80]. It also regulates chloroplast membrane fluidity [42], thereby reducing high light- or heat-induced inhibition of electron transport and limiting the degradation of key photosynthetic components [41]. Collectively, these effects contribute to improved energy conversion efficiency and sustained photosynthetic activity under abiotic stress.
Maintenance of carbon–nitrogen (C/N) metabolic balance is another crucial aspect of stress adaptation [81]. Under normal conditions, carbon assimilation (photosynthesis) provides sugars and energy, while nitrogen assimilation (nitrate uptake and assimilation) generates amino acids and proteins. Abiotic stresses often disrupt this balance by suppressing photosynthesis and nitrogen metabolism, leading to ROS buildup and metabolic imbalance [82]. GB helps reestablish C/N balance by enhancing key carbon and nitrogen metabolic enzymes. For example, under moderate drought conditions, foliar application of 25, 50, and 100 mM GB increased the soluble sugar content in sweet potato leaf tissues by 1.98-, 2.21-, and 2.37-fold, respectively, compared to the control [70]. The accumulated sugars (from carbon metabolism) and proline (from nitrogen metabolism) act as osmolytes and energy stores [83].
In addition to facilitating metabolic adjustment, GB modulates the synthesis of secondary metabolites that enhance plant stress defenses. GB has been shown to induce the biosynthesis of flavonoids and lignin, two key compounds involved in abiotic and biotic stress tolerance. Flavonoids function as potent antioxidants that scavenge ROS, while lignin reinforces cell walls, improving mechanical strength and pathogen resistance [84]. Interestingly, genes involved in GB biosynthesis share structural or regulatory features with those governing phenylpropanoid metabolism, including lignin and other phenolic compound pathways [85,86]. This suggests a possible metabolic or transcriptional linkage, whereby GB signaling may coordinate broader defensive metabolite networks. Such cross-pathway regulation further enhances antioxidant capacity and structural resilience, contributing to overall stress adaptation.

3.3. Enhancement of Plant Tolerance to Abiotic Stress by GB

3.3.1. Drought Stress

Drought is a major stress that disrupts plant water balance, reduces water-use efficiency (WUE), and compromises photosynthesis, ultimately limiting growth. It notably affects chloroplasts, where water deficit induces an over-reduction in the photosynthetic electron transport chain, resulting in ROS accumulation. This accumulation damages cellular integrity and alters the chloroplast proteome [87]. Specifically, drought stress can harm chloroplast membranes, increase ROS levels, and inhibit chlorophyll biosynthesis along with the expression of chlorophyll-binding proteins, thereby compromising PSII function [88,89].
Exogenous GB application can ameliorate drought-induced damage through multiple physiological layers. GB-treated plants often show optimized stomatal behavior, enhanced photosynthesis, stronger antioxidant defenses, and improved energy metabolism, all contributing to reduced water loss under drought. For example, in cotton (Gossypium hirsutum), Hamani et al. (2021) found that foliar GB improved stomatal conductance, gas exchange, and chlorophyll fluorescence under drought conditions [90].
At the cellular level, GB helps maintain membrane potential and water status in both the cytosol and the apoplast. It acts as an antioxidant, scavenging excess drought-induced ROS and preventing membrane damage [91]. Additionally, GB supports the photosynthetic machinery: it promotes turnover of the PSII D1 protein, alleviating photodamage and sustaining chlorophyll content and photosynthetic efficiency [92].
At the molecular level, GB influences gene expression related to cell wall and energy metabolism. Bai et al. (2022) reported that GB upregulates Cellulose Synthase A (CesA) to enhance cellulose biosynthesis, which may improve water retention. GB also modulated the expression of enzymes in energy pathways, diacylglycerol acyltransferase (DGAT) and glycerol-3-phosphate dehydrogenase (GPDH), and affected endogenous abscisic acid (ABA) levels via genes such as phospholipase D (PLD) and glutamyl-tRNA synthetase (GluRS) [93]. These changes suggest that GB coordinates osmotic signaling with metabolic adjustments under drought. In another example, applying 20 mM GB to Indian mustard (Brassica juncea) markedly improved drought resilience [94]. GB-treated mustard plants exhibited higher growth, biomass, leaf water content, nutrient uptake, and photosynthetic performance compared to controls. They also accumulated more osmolytes, exhibited stronger antioxidant enzyme activity, maintained better ion balance, and had higher cell viability under water stress. This comprehensive response confirms that GB mitigates drought stress via coordinated physiological and molecular changes.

3.3.2. Salt Stress

Salt stress poses multiple challenges to plants. Excess external Na+ disturbs cellular ion homeostasis by competing with K+ uptake and directly inhibits enzymes and protein synthesis. These disruptions lead to reduced growth, lower biomass, and yield losses. To cope, plants minimize water loss, enhance water uptake, sequester Na+ in vacuoles, or actively extrude Na+ to alleviate toxicity [95].
GB aids salt stress tolerance primarily by acting as an osmoprotectant that maintains ionic and osmotic balance [73]. Exogenous GB treatments have been shown to improve biomass and growth under salinity by increasing WUE, raising the K+/Na+ ratio, enhancing photosynthesis, bolstering ROS scavenging, and inducing stress-related genes [96,97]. For example, in maize, GB-treated plants upregulated the plasma membrane H+-ATPase, which helps extrude Na+ and thus mitigates ion toxicity [97]. In tobacco, GB reduced salt-induced oxidative damage by lowering protein carbonylation levels [98]. In Stevia (Stevia rebaudiana), GB application under salt stress improved growth by modulating nitrogen metabolism, polyamine oxidase activity, and antioxidant defenses while maintaining K+/Na+ homeostasis [99]. Intriguingly, GB also increased the expression of enzymes in the steviol glycoside biosynthesis pathway, leading to higher levels of rebaudioside A (a major sweet compound) under salinity. This suggests GB might induce protective secondary metabolites in response to salt stress. Together, these studies show that GB contributes to salt tolerance through multiple layers: it preserves osmotic balance across membranes and simultaneously upregulates antioxidant systems to prevent oxidative injury. These combined effects sustain growth and photosynthesis in saline environments.
Recent advances have expanded GB applications via nanotechnology. Ganjavi et al. (2021) created glycine betaine-loaded graphene oxide (GO-GB) nanoparticles and demonstrated that they promoted growth and mitigated salt stress in sweet basil [100]. Similarly, Hanif and Zia (2023) found that combined application of GB and zinc oxide nanoparticles (ZnO NPs) in plants significantly reduced Na+ accumulation and increased K+ levels. This combination improved ionic homeostasis and lessened salt damage, suggesting that nano-formulations can enhance GB’s efficacy in stress mitigation [101].

3.3.3. Temperature Stress

Cold stress induces ice crystal formation in plant tissues, disrupting cellular functions. Foliar application of GB acts as a cryoprotectant, enhancing cold tolerance by upregulating antioxidant defenses, protecting photosynthesis, and modulating hormone and cold-responsive pathways [102]. For example, in tomato seedlings, GB treatment increased catalase1 (CAT1) expression and activity, reduced ROS levels, preserved PSII function, and improved photosynthetic efficiency under chilling conditions [64]. Regarding hormonal balance, GB elevates ABA while suppressing GA, thereby enhancing seed cold tolerance and germination [103]. Notably, maize studies reveal a feedback loop where endogenous ABA promotes GB accumulation by activating BADH [104].
GB also activates cold-responsive transcription factors: it induces SlCBF and ICE1 in tomato, which regulate downstream genes for cold adaptation [102]. Under cold storage of fruits, GB boosts antioxidant defenses. In Chinese pear (Pyrus ussuriensis ‘Nanguo’), GB application elevated activities and transcripts of APX, CAT, and SOD, thereby reducing lipid peroxidation caused by chilling and improving fruit quality [105]. It has also been shown to modulate the balance of ABA and GA during seed dormancy, helping seeds overcome early cold stress [103]. Thus, GB’s main role in cold stress appears to be the upregulation of antioxidant systems, coupled with hormonal adjustments, which collectively protect cells from chilling injury.
Heat stress (high temperature) likewise harms plants by denaturing proteins, damaging membranes, and inhibiting photosynthesis [106,107]. GB contributes to thermotolerance through several strategies. It can induce heat shock proteins (HSPs), especially HSP70, which function as molecular chaperones during heat stress [108]. In transgenic tomato, overexpression of CODA (a choline oxidase gene from Arthrobacter globiformis) led to higher GB levels and greater heat resistance than overexpressing spinach BADH alone [109]. These CODA plants maintained higher antioxidant enzyme activities, exhibited lower ROS accumulation, and preserved PSII function under heat. They also sustained higher CO2 assimilation and PSII efficiency compared to controls. In these plants, H2O2 may act as a signaling molecule to activate heat stress genes, further enhancing thermotolerance [109]. Foliar GB application in various crops has similarly been shown to boost CAT and SOD activities under heat, reinforcing antioxidant defenses.
In summary, GB provides comprehensive protection against temperature extremes. Under heat, it helps maintain osmotic balance, stimulates HSP accumulation, and activates antioxidant systems, collectively safeguarding cellular homeostasis. Under cold, it primarily fortifies antioxidant defenses and stress signaling pathways to prevent chilling damage. Besides temperature extremes, GB has beneficial effects against other stresses such as heavy metal toxicity. Exogenous GB has been reported to alleviate heavy metal stress (e.g., chromium, cadmium) by improving growth, photosynthesis, and antioxidant enzyme activities, while reducing metal accumulation and oxidative stress in plants [110,111]. These findings suggest that GB’s protective mechanisms extend to a broad range of abiotic challenges.

3.4. Enhancing Plant Resistance to Biotic Stress

GB also contributes to biotic stress resistance by interacting with salicylic acid (SA) signaling, a key regulator of systemic acquired resistance (SAR) and pathogenesis-related (PR) protein expression. GB can enhance SA signal transduction, upregulates SA biosynthetic genes such as ICS1, and activates the SA regulator NPR1, thereby promoting PR gene expression and broad-spectrum pathogen resistance [112]. Additionally, GB helps maintain SA pathway homeostasis, preventing excessive defense activation that could hinder growth. Through these synergistic effects, GB-treated plants can more efficiently resist various pathogens, including bacteria and viruses. Additionally, GB has been demonstrated nematocidal activity. Khanna et al. (2019) reported that GB can disrupt nematode cell membranes and metabolism, providing a novel approach to managing nematode-related plant diseases [113,114].
Moreover, GB influences the production of defensive compounds. In mango and tobacco, GB-mediated defense responses involved the transcription factor MiWRKY53, which enhanced chalcone isomerase (CHI) activity, a key step in flavonoid biosynthesis. Increased CHI activity correlated with reduced disease symptoms and better chlorophyll retention in infected tissues [115]. In a groundbreaking study, foliar application of GB together with chitosan in cucumber plants greatly reduced symptoms and viral loads of cucumber mosaic virus (CMV) [112]. These results highlight GB’s potential as an inducer of disease resistance and a modulator of plant defense metabolism.

4. Strategies for Applying GB to Enhance Plant Stress Resilience

Enhancing GB levels in plants has proven to be an effective strategy for improving plant stress tolerance. Two main approaches are commonly employed: exogenous application and genetic engineering of GB biosynthetic pathways (Figure 2). Both methods increase GB accumulation, activate stress-responsive pathways, and enhance resilience to a range of abiotic stresses.

4.1. Exogenous Application of GB

The exogenous application of GB, through foliar spray, seed treatment, or soil drenching, is a practical strategy to enhance plant stress tolerance. Among these, foliar application is especially effective, as GB is efficiently absorbed through leaves, leading to a rapid increase in internal concentration (Table 1). For example, treating peach (Prunus persica) with GB induced changes in cellular energy status, raised internal levels of GB, γ-aminobutyric acid (GABA), and proline, and protected membranes, collectively enhancing cold tolerance [116]. In another case, spraying GB on licorice (Glycyrrhiza uralensis) seedlings under salt stress stimulated osmolyte accumulation, elevated antioxidant enzyme activities, promoted salt excretion, and significantly increased Na+ efflux, thereby improving salt tolerance [117]. Similarly, applying 30 mM GB to Chinese cabbage leaves improved freezing tolerance by reducing membrane damage and ROS accumulation without harming plant health [118].
Exogenous GB acts through multiple regulatory mechanisms. In maize, Bai et al. (2022) showed that GB alleviated drought stress by modulating endogenous ABA levels and stomatal conductance via stress-responsive gene expression [93]. In Arabidopsis, Zhang et al. (2022) identified the gene PCST1 as mediating osmotic stress responses by regulating proline and GB biosynthesis, as well as SOS, NCED, and CIPK genes [146]. In tomato, GB was found to mitigate cold-induced inhibition of seed germination through several signaling pathways [103]. GB application has also improved postharvest fruit quality: it enhanced flavor and cold tolerance in stored peaches [147] and extended the storability and shelf life of Citrus reticulata ‘Huangguogan’ [129].
Despite these benefits, exogenous GB has limitations. Its uptake, distribution, and metabolic fate in plants can vary with species, developmental stage, and environmental conditions, leading to inconsistent stress mitigation. Large-scale application of GB can also be costly and may pose ecological risks, such as unintended effects on soil microbiota and surrounding ecosystems. Careful optimization of dosage, timing, and delivery methods is therefore necessary for practical use.

4.2. Genetic Engineering

Genetic engineering provides an alternative strategy to enhance GB accumulation by introducing key biosynthetic enzymes. Overexpression of GB biosynthetic genes has been shown to improve stress tolerance across various species. For example, transgenic soybean lines expressing the Atriplex hortensis BADH gene (AhBADH) exhibited significantly increased salt tolerance in transgenic lines [148]. Similarly, tobacco plants overexpressing spinach BADH demonstrated enhanced resistance to cadmium stress [149].
Interestingly, enzymes in the GB pathway can also impact crop quality traits. In rice, the BADH2 gene is closely linked to fragrance: mutations that reduce BADH2 activity increase the aroma compound 2-acetyl-1-pyrroline (2AP) [22,150]. Recent studies have mapped expression and protein QTLs for BADH2, revealing its major role in 2AP variation [151]. Novel BADH2 alleles are being used to breed fragrant rice (Hui et al., 2022). In peanut (Arachis hypogaea), CRISPR/Cas9 knockout of AhBADH1 and AhBADH2 produced lines with dramatically higher 2AP content and enhanced aroma [152]. These examples show that the manipulation of GB biosynthesis genes can have broader metabolic consequences, affecting secondary traits like flavor.
However, genetic modifications targeting aroma-related genes may also impact plant responses to abiotic stress. For example, under increasing salinity, both the transcript levels of BADH1 and the BADH1/BADH2 transcript ratio were significantly elevated in the leaf tissues of both aromatic and non-aromatic rice cultivars. In contrast, BADH2 transcript levels displayed no consistent pattern in response to salt stress. These observations suggest that BADH1, rather than BADH2, plays a functional role in mediating salt stress responses in rice [153]. While targeted knockout of the OsBADH2 gene induces aroma formation, phenotypic responses to salinity vary depending on the genetic background. For instance, CRISPR/Cas9-induced mutants of two non-aromatic rice varieties, Huaidao#5 and Jiahua#1, both developed aromas but exhibited opposite responses to salt stress: enhanced tolerance in Huaidao#5 and reduced tolerance in Jiahua#1 [154]. Physiological and transcriptional analyses indicated differential changes in osmolyte levels and stress-responsive and antioxidant enzymatic gene expression, suggesting that interactions between genetic background and OsBADH2 disruption play a critical role in determining stress response outcomes. Therefore, further molecular investigations are essential to elucidate the regulatory mechanisms linking OsBADH gene function with both aroma biosynthesis and abiotic stress tolerance in rice.
Despite the engineering of GB pathways showing promise, challenges remain. Transgene expression must be stable and robust; factors like genomic integration site, promoter choice, and epigenetics can affect performance. Overexpressing GB biosynthetic enzymes may also disrupt metabolic homeostasis, potentially impairing growth. Therefore, future efforts should focus on understanding the regulatory networks governing GB metabolism and on finding effective tissue-specific or stress-inducible promoters. Optimizing transformation protocols and coordinating GB gene expression with the plant’s overall metabolism will be essential. Only by balancing enhanced stress tolerance with normal development can these approaches achieve practical agricultural benefits.

5. Plant GB in Agriculture: Mechanisms, Challenges, and Future Perspectives

As a natural osmoprotectant, GB holds substantial promise for improving crop stress resistance. Whether through endogenous synthesis or exogenous application, GB significantly enhances plant tolerance to a broad spectrum of abiotic and biotic stresses. Elucidating its molecular mechanisms of action is thus a key focus in plant science and crop improvement.
GB is predominantly synthesized via the choline oxidation pathway in plants or via the glycine methylation pathway in certain microbes. The capacity for GB accumulation varies widely among species and is influenced by environmental conditions and developmental stage. Mechanistically, GB maintains cellular osmotic homeostasis and stabilizes membranes and proteins, mitigating water loss and cellular damage under drought, salinity, or temperature extremes. Its antioxidant function is especially noteworthy: GB both directly neutralizes ROS and upregulates enzymes like SOD and CAT, thereby alleviating oxidative cytotoxicity. Additionally, GB modulates carbon-nitrogen metabolism by promoting osmolyte synthesis (e.g., proline, sugars) and sustaining energy supply under stress. Extensive studies have also shown that GB participates in stress signaling. It influences hormone signaling pathways (e.g., ABA and SA) and activates stress-responsive transcriptional networks under drought, salt, and cold. These signaling roles further integrate GB into the plant’s defense responses.
In terms of applications, exogenous GB supply is a simple and effective way to improve plant performance under stress. However, its efficacy depends on uptake and transport efficiency, which vary with plant species, tissue type, and application method. Consequently, stress protection by GB can be inconsistent. Transgenic strategies, in contrast, have achieved promising results: introducing CMO and BADH genes has enhanced stress tolerance in many crops. Intriguingly, GB biosynthesis genes can also affect quality traits like grain aroma in rice. Yet genetic approaches still face hurdles, such as unstable transgene expression and possible metabolic imbalances. Notably, the mechanisms of GB uptake and transport across plant membranes (e.g., potential H+-coupled GB transporters) remain poorly understood, representing a bottleneck for both basic research and practical deployment.
Future research should focus on unraveling the molecular networks regulating GB biosynthesis, signaling, and transport, particularly its crosstalk with plant hormones, transcription factors, and redox regulators. Integrating advanced approaches such as gene editing, synthetic biology, metabolic engineering, and nano-based delivery systems will facilitate precise manipulation of GB accumulation and activity. For example, the development of eco-friendly nano-formulations may enhance the efficiency of GB delivery under field conditions. Despite existing challenges in regulation, application efficiency, and field translation, GB remains a promising metabolic tool for enhancing stress adaptation. Its potential to support stress-resilient crop breeding and promote sustainable agriculture is especially valuable in the context of global climate change and agroecosystem stability [155].

Author Contributions

Conceptualization, J.C., J.Z. and Y.X.; software, J.C., K.Z. and Y.L.; writing—original draft preparation, J.C. and J.Z.; writing—review and editing, K.Z. and Y.L.; supervision, F.Z. and Y.X.; project administration, F.Z. and Y.X.; funding acquisition, F.Z. and Y.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work is supported by grants from the Natural Science Foundation of China (32470299), the Jiangsu Natural Science Foundation for Distinguished Young Scholars (BK20220084), and the STI 2030-Major Projects (2023ZD040562).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, P.; Liu, W.C.; Han, C.; Wang, S.; Bai, M.Y.; Song, C.P. Reactive oxygen species: Multidimensional regulators of plant adaptation to abiotic stress and development. J. Integr. Plant Biol. 2024, 66, 330–367. [Google Scholar] [CrossRef]
  2. Zhu, J.K. Salt and drought stress signal transduction in plants. Annu. Rev. Plant Biol. 2002, 53, 247–273. [Google Scholar] [CrossRef]
  3. Rhodes, D.; Hanson, A. Quaternary ammonium and tertiary sulfonium compounds in higher plants. Annu. Rev. Plant Biol. 1993, 44, 357–384. [Google Scholar] [CrossRef]
  4. Hasanuzzaman, M.; Banerjee, A.; Bhuyan, M.-H.-M.-B.; Roychoudhury, A.; Mahmud, J.-A. Targeting Glycinebetaine for Abiotic Stress Tolerance in Crop Plants: Physiological Mechanism, Molecular Interaction and Signaling. Phyton-Int. J. Exp. Bot. 2019, 88, 185–221. [Google Scholar] [CrossRef]
  5. Lutts, S. Exogenous glycinebetaine reduces sodium accumulation in salt-stressed rice plants. Int. Rice Res. Notes 2000, 25, 39–40. [Google Scholar] [CrossRef]
  6. Rathinasabapathi, B.; Burnet, M.; Russell, B.L.; Gage, D.A.; Liao, P.C.; Nye, G.J.; Scott, P.; Golbeck, J.H.; Hanson, A.D. Choline monooxygenase, an unusual iron-sulfur enzyme catalyzing the first step of glycine betaine synthesis in plants: Prosthetic group characterization and cDNA cloning. Proc. Natl. Acad. Sci. USA 1997, 94, 3454–3458. [Google Scholar] [CrossRef] [PubMed]
  7. Yamada, N.; Takahashi, H.; Kitou, K.; Sahashi, K.; Tamagake, H.; Tanaka, Y.; Takabe, T. Suppressed expression of choline monooxygenase in sugar beet on the accumulation of glycine betaine. Plant Physiol. Biochem. 2015, 96, 217–221. [Google Scholar] [CrossRef] [PubMed]
  8. Russell, B.L.; Rathinasabapathi, B.; Hanson, A.D. Osmotic Stress Induces Expression of Choline Monooxygenase in Sugar Beet and Amaranth. Plant Physiol. 1998, 116, 859–865. [Google Scholar] [CrossRef]
  9. Hanke, G.; Mulo, P. Plant type ferredoxins and ferredoxin-dependent metabolism. Plant Cell Environ. 2013, 36, 1071–1084. [Google Scholar] [CrossRef]
  10. Carrillo, N.; Ceccarelli, E.A. Open questions in ferredoxin-NADP+ reductase catalytic mechanism. Eur. J. Biochem. 2003, 270, 1900–1915. [Google Scholar] [CrossRef]
  11. Chauhan, J.; Prathibha, M.D.; Singh, P.; Choyal, P.; Mishra, U.N.; Saha, D.; Kumar, R.; Anuragi, H.; Pandey, S.; Bose, B.; et al. Plant photosynthesis under abiotic stresses: Damages, adaptive, and signaling mechanisms. Plant Stress. 2023, 10, 100296. [Google Scholar] [CrossRef]
  12. Weigel, P.; Weretilnyk, E.A.; Hanson, A.D. Betaine aldehyde oxidation by spinach chloroplasts. Plant Physiol. 1986, 82, 753–759. [Google Scholar] [CrossRef] [PubMed]
  13. Nakamura, T.; Yokota, S.; Muramoto, Y.; Tsutsui, K.; Oguri, Y.; Fukui, K.; Takabe, T. Expression of a betaine aldehyde dehydrogenase gene in rice, a glycinebetaine nonaccumulator, and possible localization of its protein in peroxisomes. Plant J. 1997, 11, 1115–1120. [Google Scholar] [CrossRef] [PubMed]
  14. Pan, S.M.; Moreau, R.A.; Yu, C.; Huang, A.H. Betaine accumulation and betaine-aldehyde dehydrogenase in spinach leaves. Plant Physiol. 1981, 67, 1105–1108. [Google Scholar] [CrossRef] [PubMed]
  15. Arakawa, K.; Katayama, M.; Takabe, T. Levels of Betaine and Betaine Aldehyde Dehydrogenase Activity in the Green Leaves, and Etiolated Leaves and Roots of Barley1. Plant Cell Physiol. 1990, 31, 797–803. [Google Scholar] [CrossRef]
  16. Li, P.; Cai, J.; Luo, X.; Chang, T.; Li, J.; Zhao, Y.; Xu, Y. Transformation of wheat Triticum aestivum with the HvBADH1 transgene from hulless barley improves salinity-stress tolerance. Acta Physiologiae Plantarum 2019, 41, 155. [Google Scholar] [CrossRef]
  17. Hasthanasombut, S.; Supaibulwatana, K.; Mii, M.; Nakamura, I. Genetic manipulation of Japonica rice using the OsBADH1 gene from Indica rice to improve salinity tolerance. Plant Cell Tissue Org. Culture 2011, 104, 79–89. [Google Scholar] [CrossRef]
  18. Breisch, J.; Bendel, M.; Averhoff, B. The choline dehydrogenase BetA of Acinetobacter baumannii: A flavoprotein responsible for osmotic stress protection. Environ. Microbiol. 2022, 24, 1052–1061. [Google Scholar] [CrossRef]
  19. Huang, H.; Song, D.; Zhang, W.; Fang, S.; Zhou, Q.; Zhang, H.; Liang, Z.; Li, Y. Choline Oxidase-Integrated Copper Metal-Organic Frameworks as Cascade Nanozymes for One-Step Colorimetric Choline Detection. J. Agric. Food Chem. 2022, 70, 5228–5236. [Google Scholar] [CrossRef]
  20. Shirasawa, K.; Takabe, T.; Takabe, T.; Kishitani, S. Accumulation of glycinebetaine in rice plants that overexpress choline monooxygenase from spinach and evaluation of their tolerance to abiotic stress. Ann. Bot. 2006, 98, 565–571. [Google Scholar] [CrossRef]
  21. Mitsuya, S.; Yokota, Y.; Fujiwara, T.; Mori, N.; Takabe, T. OsBADH1 is possibly involved in acetaldehyde oxidation in rice plant peroxisomes. FEBS Lett. 2009, 583, 3625–3629. [Google Scholar] [CrossRef]
  22. Niu, X.; Tang, W.; Huang, W.; Ren, G.; Wang, Q.; Luo, D.; Xiao, Y.; Yang, S.; Wang, F.; Lu, B.R.; et al. RNAi-directed downregulation of OsBADH2 results in aroma (2-acetyl-1-pyrroline) production in rice (Oryza sativa L.). BMC Plant Biol. 2008, 8, 100. [Google Scholar] [CrossRef]
  23. Missihoun, T.D.; Willee, E.; Guegan, J.P.; Berardocco, S.; Shafiq, M.R.; Bouchereau, A.; Bartels, D. Overexpression of ALDH10A8 and ALDH10A9 Genes Provides Insight into Their Role in Glycine Betaine Synthesis and Affects Primary Metabolism in Arabidopsis thaliana. Plant Cell Physiol. 2015, 56, 1798–1807. [Google Scholar] [CrossRef] [PubMed]
  24. Huang, J.; Hirji, R.; Adam, L.; Rozwadowski, K.L.; Hammerlindl, J.K.; Keller, W.A.; Selvaraj, G. Genetic engineering of glycinebetaine production toward enhancing stress tolerance in plants: Metabolic limitations. Plant Physiol. 2000, 122, 747–756. [Google Scholar] [CrossRef] [PubMed]
  25. Hibino, T.; Waditee, R.; Araki, E.; Ishikawa, H.; Aoki, K.; Tanaka, Y.; Takabe, T. Functional characterization of choline monooxygenase, an enzyme for betaine synthesis in plants. J. Biol. Chem. 2002, 277, 41352–41360. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, J.; Tan, W.; Yang, X.-H.; Zhang, H.-X. Plastid-expressed choline monooxygenase gene improves salt and drought tolerance through accumulation of glycine betaine in tobacco. Plant Cell Rep. 2008, 27, 1113–1124. [Google Scholar] [CrossRef]
  27. Nuccio, M.L.; Russell, B.L.; Nolte, K.D.; Rathinasabapathi, B.; Gage, D.A.; Hanson, A.D. The endogenous choline supply limits glycine betaine synthesis in transgenic tobacco expressing choline monooxygenase. Plant J. 1998, 16, 487–496. [Google Scholar] [CrossRef]
  28. Hayashi, H.; Alia; Mustardy, L.; Deshnium, P.; Ida, M.; Murata, N. Transformation of Arabidopsis thaliana with the codA gene for choline oxidase; accumulation of glycinebetaine and enhanced tolerance to salt and cold stress. Plant J. 1997, 12, 133–142. [Google Scholar] [CrossRef]
  29. Yilmaz, J.L.; Bulow, L. Enhanced stress tolerance in Escherichia coli and Nicotiana tabacum expressing a betaine aldehyde dehydrogenase/choline dehydrogenase fusion protein. Biotechnol. Prog. 2002, 18, 1176–1182. [Google Scholar] [CrossRef]
  30. Cheng, Y.J.; Deng, X.P.; Kwak, S.S.; Chen, W.; Eneji, A.E. Enhanced tolerance of transgenic potato plants expressing choline oxidase in chloroplasts against water stress. Bot. Stud. 2013, 54, 30. [Google Scholar] [CrossRef]
  31. Li, Q.; Yin, H.; Li, D.; Zhu, H.; Zhang, Y.; Zhu, W. Isolation and characterization of CMO gene promoter from halophyte Suaeda liaotungensis K. J. Genet. Genom. 2007, 34, 355–361. [Google Scholar] [CrossRef] [PubMed]
  32. Lescot, M.; Dehais, P.; Thijs, G.; Marchal, K.; Moreau, Y.; Van de Peer, Y.; Rouze, P.; Rombauts, S. PlantCARE, a database of plant cis-acting regulatory elements and a portal to tools for in silico analysis of promoter sequences. Nucleic Acids Res. 2002, 30, 325–327. [Google Scholar] [CrossRef] [PubMed]
  33. Maruyama, K.; Todaka, D.; Mizoi, J.; Yoshida, T.; Kidokoro, S.; Matsukura, S.; Takasaki, H.; Sakurai, T.; Yamamoto, Y.Y.; Yoshiwara, K.; et al. Identification of cis-acting promoter elements in cold- and dehydration-induced transcriptional pathways in Arabidopsis, rice, and soybean. DNA Res. 2012, 19, 37–49. [Google Scholar] [CrossRef] [PubMed]
  34. Fan, X.; Peng, L.; Zhang, Y. Plant DNA methylation responds to nutrient stress. Genes. 2022, 13, 992. [Google Scholar] [CrossRef]
  35. Liu, Y.; Wang, J.; Liu, B.; Xu, Z.Y. Dynamic regulation of DNA methylation and histone modifications in response to abiotic stresses in plants. J. Integr. Plant Biol. 2022, 64, 2252–2274. [Google Scholar] [CrossRef]
  36. Nunez-Vazquez, R.; Desvoyes, B.; Gutierrez, C. Histone variants and modifications during abiotic stress response. Front. Plant Sci. 2022, 13, 984702. [Google Scholar] [CrossRef]
  37. Sun, M.; Yang, Z.; Liu, L.; Duan, L. DNA methylation in plant responses and adaption to abiotic stresses. Int. J. Mol. Sci. 2022, 23, 6910. [Google Scholar] [CrossRef]
  38. Zi, N.; Ren, W.; Guo, H.; Yuan, F.; Liu, Y.; Fry, E. DNA Methylation Participates in Drought Stress Memory and Response to Drought in Medicago ruthenica. Genes 2024, 15, 1286. [Google Scholar] [CrossRef]
  39. McNeil, S.D.; Nuccio, M.L.; Ziemak, M.J.; Hanson, A.D. Enhanced synthesis of choline and glycine betaine in transgenic tobacco plants that overexpress phosphoethanolamine N-methyltransferase. Proc. Natl. Acad. Sci. USA 2001, 98, 10001–10005. [Google Scholar] [CrossRef]
  40. Hanson, A.D.; Rhodes, D. C Tracer Evidence for Synthesis of Choline and Betaine via Phosphoryl Base Intermediates in Salinized Sugarbeet Leaves. Plant Physiol. 1983, 71, 692–700. [Google Scholar] [CrossRef]
  41. Papageorgiou, G.C.; Murata, N. The unusually strong stabilizing effects of glycine betaine on the structure and function of the oxygen-evolving photosystem II complex. Photosynth. Res. 1995, 44, 243–252. [Google Scholar] [CrossRef]
  42. Chen, T.H.; Murata, N. Glycinebetaine protects plants against abiotic stress: Mechanisms and biotechnological applications. Plant Cell Environ. 2011, 34, 1–20. [Google Scholar] [CrossRef] [PubMed]
  43. Wood, A.J.; Saneoka, H.; Rhodes, D.; Joly, R.J.; Goldsbrough, P.B. Betaine aldehyde dehydrogenase in sorghum. Plant Physiol. 1996, 110, 1301–1308. [Google Scholar] [CrossRef] [PubMed]
  44. Zhang, Y.; Yin, H.; Li, D.; Zhu, W.; Li, Q. Functional analysis of BADH gene promoter from Suaeda liaotungensis K. Plant Cell Rep. 2008, 27, 585–592. [Google Scholar] [CrossRef] [PubMed]
  45. Nyyssola, A.; Kerovuo, J.; Kaukinen, P.; von Weymarn, N.; Reinikainen, T. Extreme halophiles synthesize betaine from glycine by methylation. J. Biol. Chem. 2000, 275, 22196–22201. [Google Scholar] [CrossRef]
  46. Waditee, R.; Tanaka, Y.; Aoki, K.; Hibino, T.; Jikuya, H.; Takano, J.; Takabe, T.; Takabe, T. Isolation and functional characterization of N-methyltransferases that catalyze betaine synthesis from glycine in a halotolerant photosynthetic organism Aphanothece halophytica. J. Biol. Chem. 2003, 278, 4932–4942. [Google Scholar] [CrossRef]
  47. Nyyssola, A.; Reinikainen, T.; Leisola, M. Characterization of glycine sarcosine N-methyltransferase and sarcosine dimethylglycine N-methyltransferase. Appl. Environ. Microbiol. 2001, 67, 2044–2050. [Google Scholar] [CrossRef]
  48. Waditee, R.; Bhuiyan, M.N.; Rai, V.; Aoki, K.; Tanaka, Y.; Hibino, T.; Suzuki, S.; Takano, J.; Jagendorf, A.T.; Takabe, T.; et al. Genes for direct methylation of glycine provide high levels of glycinebetaine and abiotic-stress tolerance in Synechococcus and Arabidopsis. Proc. Natl. Acad. Sci. USA 2005, 102, 1318–1323. [Google Scholar] [CrossRef]
  49. Lai, S.J.; Lai, M.C.; Lee, R.J.; Chen, Y.H.; Yen, H.E. Transgenic Arabidopsis expressing osmolyte glycine betaine synthesizing enzymes from halophilic methanogen promote tolerance to drought and salt stress. Plant Mol. Biol. 2014, 85, 429–441. [Google Scholar] [CrossRef]
  50. Rajan, L.A.; Joseph, T.C.; Thampuran, N.; James, R. Functional characterization and sequence analysis of choline dehydrogenase from Escherichia coli. Genet. Eng. Biotechnol. J. 2010, 2010, GEBJ-12. [Google Scholar]
  51. Zhao, F.; Luo, J.; Ibrahim, E.; Chen, L.; Shen, Y.; Ibrahim, M.; Alonazi, W.B.; Lu, J.; Luo, Y.; Wu, H. Engineering stress resistance: Advances in glycine betaine production for sustainable agriculture. Crop Health 2025, 3, 5. [Google Scholar] [CrossRef]
  52. Kleczkowski, L.A.; Givan, C.V. Serine formation in leaves by mechanisms other than the glycolate pathway. J. Plant Physiol. 1988, 132, 641–652. [Google Scholar] [CrossRef]
  53. Igamberdiev, A.U.; Kleczkowski, L.A. The glycerate and phosphorylated pathways of serine synthesis in plants: The branches of plant glycolysis linking carbon and nitrogen metabolism. Front. Plant Sci. 2018, 9, 318. [Google Scholar] [CrossRef] [PubMed]
  54. Wingler, A.; Lea, P.J.; Quick, W.P.; Leegood, R.C. Photorespiration: Metabolic pathways and their role in stress protection. Philos. Trans. R. Soc. B Biol. Sci. 2000, 355, 1517–1529. [Google Scholar] [CrossRef] [PubMed]
  55. Carillo, P.; Mastrolonardo, G.; Nacca, F.; Parisi, D.; Verlotta, A.; Fuggi, A. Nitrogen metabolism in durum wheat under salinity: Accumulation of proline and glycine betaine. Funct. Plant Biol. 2008, 35, 412–426. [Google Scholar] [CrossRef]
  56. Annunziata, M.G.; Ciarmiello, L.F.; Woodrow, P.; Dell’Aversana, E.; Carillo, P. Spatial and Temporal Profile of Glycine Betaine Accumulation in Plants Under Abiotic Stresses. Front. Plant Sci. 2019, 10, 230. [Google Scholar] [CrossRef]
  57. Michel, V.; Yuan, Z.; Ramsubir, S.; Bakovic, M. Choline transport for phospholipid synthesis. Exp. Biol. Med. 2006, 231, 490–504. [Google Scholar] [CrossRef]
  58. Dettmer, J.; Ursache, R.; Campilho, A.; Miyashima, S.; Belevich, I.; O’Regan, S.; Mullendore, D.L.; Yadav, S.R.; Lanz, C.; Beverina, L.; et al. CHOLINE TRANSPORTER-LIKE1 is required for sieve plate development to mediate long-distance cell-to-cell communication. Nat. Commun. 2014, 5, 4276. [Google Scholar] [CrossRef]
  59. McNeil, S.D.; Rhodes, D.; Russell, B.L.; Nuccio, M.L.; Shachar-Hill, Y.; Hanson, A.D. Metabolic modeling identifies key constraints on an engineered glycine betaine synthesis pathway in tobacco. Plant Physiol. 2000, 124, 153–162. [Google Scholar] [CrossRef]
  60. Khan, M.S.; Yu, X.; Kikuchi, A.; Asahina, M.; Watanabe, K.N. Genetic engineering of glycine betaine biosynthesis to enhance abiotic stress tolerance in plants. Plant Biotechnol. 2009, 26, 125–134. [Google Scholar] [CrossRef]
  61. Yamada, N.; Sakakibara, S.; Tsutsumi, K.; Waditee, R.; Tanaka, Y.; Takabe, T. Expression and substrate specificity of betaine/proline transporters suggest a novel choline transport mechanism in sugar beet. J. Plant Physiol. 2011, 168, 1609–1616. [Google Scholar] [CrossRef]
  62. Yang, T.; Nian, Y.; Lin, H.; Li, J.; Lin, X.; Li, T.; Wang, R.; Wang, L.; Beattie, G.A.; Zhang, J.; et al. Structure and mechanism of the osmoregulated choline transporter BetT. Sci. Adv. 2024, 10, eado6229. [Google Scholar] [CrossRef]
  63. Kumar, V.; Shriram, V.; Hoque, T.S.; Hasan, M.M.; Burritt, D.J.; Hossain, M.A. Glycinebetaine-mediated abiotic oxidative-stress tolerance in plants: Physiological and biochemical mechanisms. Stress. Signal. Plants Genom. Proteom. Perspect. 2017, 2, 111–133. [Google Scholar] [CrossRef]
  64. Park, E.-J.; Jeknic, Z.; Chen, T.H. Exogenous application of glycinebetaine increases chilling tolerance in tomato plants. Plant Cell Physiol. 2006, 47, 706–714. [Google Scholar] [CrossRef] [PubMed]
  65. Sulpice, R.; Tsukaya, H.; Nonaka, H.; Mustardy, L.; Chen, T.H.; Murata, N. Enhanced formation of flowers in salt-stressed Arabidopsis after genetic engineering of the synthesis of glycine betaine. Plant J. 2003, 36, 165–176. [Google Scholar] [CrossRef] [PubMed]
  66. Park, E.J.; Jeknic, Z.; PINO, M.T.; Murata, N.; CHEN, T.H.H. Glycinebetaine accumulation is more effective in chloroplasts than in the cytosol for protecting transgenic tomato plants against abiotic stress. Plant Cell Environ. 2007, 30, 994–1005. [Google Scholar] [CrossRef] [PubMed]
  67. Ozturk, M.; Turkyilmaz Unal, B.; García-Caparros, P.; Khursheed, A.; Gul, A.; Hasanuzzaman, M. Osmoregulation and its actions during the drought stress in plants. Physiol. Plant 2021, 172, 1321–1335. [Google Scholar] [CrossRef]
  68. Xu, Z.; Sun, M.; Jiang, X.; Sun, H.; Dang, X.; Cong, H.; Qiao, F. Glycinebetaine Biosynthesis in Response to Osmotic Stress Depends on Jasmonate Signaling in Watermelon Suspension Cells. Front. Plant Sci. 2018, 9, 1469. [Google Scholar] [CrossRef]
  69. Wani, S.H.; Singh, N.B.; Haribhushan, A.; Mir, J.I. Compatible solute engineering in plants for abiotic stress tolerance-role of glycine betaine. Curr. Genom. 2013, 14, 157–165. [Google Scholar] [CrossRef]
  70. Tisarum, R.; Theerawitaya, C.; Samphumphuang, T.; Singh, H.P.; Cha-Um, S. Foliar application of glycinebetaine regulates soluble sugars and modulates physiological adaptations in sweet potato (Ipomoea batatas) under water deficit. Protoplasma 2020, 257, 197–211. [Google Scholar] [CrossRef]
  71. 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]
  72. Quan, R.; Shang, M.; Zhang, H.; Zhao, Y.; Zhang, J. Engineering of enhanced glycine betaine synthesis improves drought tolerance in maize. Plant Biotechnol. J. 2004, 2, 477–486. [Google Scholar] [CrossRef] [PubMed]
  73. Chen, F.; Fang, P.; Zeng, W.; Ding, Y.; Zhuang, Z.; Peng, Y. Comparing transcriptome expression profiles to reveal the mechanisms of salt tolerance and exogenous glycine betaine mitigation in maize seedlings. PLoS ONE 2020, 15, e0233616. [Google Scholar] [CrossRef] [PubMed]
  74. Sies, H.; Mailloux, R.J.; Jakob, U. Fundamentals of redox regulation in biology. Nat. Rev. Mol. Cell Biol. 2024, 25, 701–719. [Google Scholar] [CrossRef]
  75. Demiral, T.; Türkan, I. Does exogenous glycinebetaine affect antioxidative system of rice seedlings under NaCl treatment? J. Plant Physiol. 2004, 161, 1089–1100. [Google Scholar] [CrossRef]
  76. Li, N.; Gao, Y.; Pu, K.; Zhang, M.; Wang, T.; Li, J.; Xie, J. Glycine betaine enhances tolerance of low temperature combined with low light in pepper (Capsicum annuum L.) by improving the antioxidant capacity and regulating GB metabolism. Plant Physiol. Biochem. 2025, 222, 109705. [Google Scholar] [CrossRef]
  77. Renne, M.F.; Ernst, R. Membrane homeostasis beyond fluidity: Control of membrane compressibility. Trends Biochem. Sci. 2023, 48, 963–977. [Google Scholar] [CrossRef]
  78. Sakamoto, A.; Murata, N. The role of glycine betaine in the protection of plants from stress: Clues from transgenic plants. Plant Cell Environ. 2002, 25, 163–171. [Google Scholar] [CrossRef]
  79. Chen, T.H.; Murata, N. Enhancement of tolerance of abiotic stress by metabolic engineering of betaines and other compatible solutes. Curr. Opin. Plant Biol. 2002, 5, 250–257. [Google Scholar] [CrossRef]
  80. Zhao, X.X.; Ma, Q.Q.; Liang, C.; Fang, Y.; Wang, Y.Q.; Wang, W. Effect of glycinebetaine on function of thylakoid membranes in wheat flag leaves under drought stress. Biol. Plantarum 2007, 51, 584–588. [Google Scholar] [CrossRef]
  81. Huang, W.; Han, S.; Wang, L.; Li, W. Carbon and nitrogen metabolic regulation in freshwater plant Ottelia alismoides in response to carbon limitation: A metabolite perspective. Front. Plant Sci. 2022, 13, 962622. [Google Scholar] [CrossRef] [PubMed]
  82. Mateo, A.; Funck, D.; Mühlenbock, P.; Kular, B.; Mullineaux, P.M.; Karpinski, S. Controlled levels of salicylic acid are required for optimal photosynthesis and redox homeostasis. J. Exp. Bot. 2006, 57, 1795–1807. [Google Scholar] [CrossRef] [PubMed]
  83. Verbruggen, N.; Hermans, C. Proline accumulation in plants: A review. Amino Acids 2008, 35, 753–759. [Google Scholar] [CrossRef] [PubMed]
  84. Dixon, R.A.; Achnine, L.; Kota, P.; Liu, C.J.; Reddy, M.S.; Wang, L. The phenylpropanoid pathway and plant defence-a genomics perspective. Mol. Plant Pathol. 2002, 3, 371–390. [Google Scholar] [CrossRef]
  85. Bhuiyan, N.H.; Selvaraj, G.; Wei, Y.; King, J. Gene expression profiling and silencing reveal that monolignol biosynthesis plays a critical role in penetration defence in wheat against powdery mildew invasion. J. Exp. Bot. 2009, 60, 509–521. [Google Scholar] [CrossRef]
  86. 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]
  87. Sha, S.; Cai, G.; Liu, S.; Ahmed, M.A. Roots to the rescue: How plants harness hydraulic redistribution to survive drought across contrasting soil textures. Adv. Biotechnol. 2024, 2, 43. [Google Scholar] [CrossRef]
  88. Chen, Y.E.; Liu, W.J.; Su, Y.Q.; Cui, J.M.; Zhang, Z.W.; Yuan, M.; Zhang, H.Y.; Yuan, S. Different response of photosystem II to short and long-term drought stress in Arabidopsis thaliana. Physiol. Plant 2016, 158, 225–235. [Google Scholar] [CrossRef]
  89. Karami, S.; Shiran, B.; Ravash, R. Molecular investigation of how drought stress affects chlorophyll metabolism and photosynthesis in leaves of C3 and C4 plant species: A transcriptome meta-analysis. Heliyon 2025, 11, e42368. [Google Scholar] [CrossRef]
  90. Hamani, A.K.M.; Li, S.; Chen, J.; Amin, A.S.; Wang, G.; Xiaojun, S.; Zain, M.; Gao, Y. Linking exogenous foliar application of glycine betaine and stomatal characteristics with salinity stress tolerance in cotton (Gossypium hirsutum L.) seedlings. BMC Plant Biol. 2021, 21, 146. [Google Scholar] [CrossRef]
  91. Gupta, P.; Rai, R.; Vasudev, S.; Yadava, D.K.; Dash, P.K. Ex-foliar application of glycine betaine and its impact on protein, carbohydrates and induction of ROS scavenging system during drought stress in flax (Linum usitatissimum). J. Biotechnol. 2021, 337, 80–89. [Google Scholar] [CrossRef]
  92. Li, D.; Wang, M.; Zhang, T.; Chen, X.; Li, C.; Liu, Y.; Brestic, M.; Chen, T.H.H.; Yang, X. Glycinebetaine mitigated the photoinhibition of photosystem II at high temperature in transgenic tomato plants. Photosynth. Res. 2021, 147, 301–315. [Google Scholar] [CrossRef] [PubMed]
  93. Bai, M.; Zeng, W.; Chen, F.; Ji, X.; Zhuang, Z.; Jin, B.; Wang, J.; Jia, L.; Peng, Y. Transcriptome expression profiles reveal response mechanisms to drought and drought-stress mitigation mechanisms by exogenous glycine betaine in maize. Biotechnol. Lett. 2022, 44, 367–386. [Google Scholar] [CrossRef] [PubMed]
  94. Islam, S.; Mohammad, F.; Shakeel, A.; Corpas, F.J. Glycine betaine: A multifaceted protectant against salt stress in Indian mustard through ionic homeostasis, ROS scavenging and osmotic regulation. Physiol. Plant 2024, 176, e14530. [Google Scholar] [CrossRef] [PubMed]
  95. Deinlein, U.; Stephan, A.B.; Horie, T.; Luo, W.; Xu, G.; Schroeder, J.I. Plant salt-tolerance mechanisms. Trends Plant Sci. 2014, 19, 371–379. [Google Scholar] [CrossRef] [PubMed]
  96. Sofy, M.R.; Elhawat, N.; Tarek, A. Glycine betaine counters salinity stress by maintaining high K+/Na+ ratio and antioxidant defense via limiting Na+ uptake in common bean (Phaseolus vulgaris L.). Ecotoxicol. Environ. Saf. 2020, 200, 110732. [Google Scholar] [CrossRef]
  97. Zhu, M.; Li, Q.; Zhang, Y.; Zhang, M.; Li, Z. Glycine betaine increases salt tolerance in maize (Zea mays L.) by regulating Na+ homeostasis. Front. Plant Sci. 2022, 13, 978304. [Google Scholar] [CrossRef]
  98. Hoque, M.A.; Banu, M.N.; Nakamura, Y.; Shimoishi, Y.; Murata, Y. Proline and glycinebetaine enhance antioxidant defense and methylglyoxal detoxification systems and reduce NaCl-induced damage in cultured tobacco cells. J. Plant Physiol. 2008, 165, 813–824. [Google Scholar] [CrossRef]
  99. Ghorbani, A.; Ghasemi-Omran, V.O.; Chen, M. The Effect of Glycine Betaine on Nitrogen and Polyamine Metabolisms, Expression of Glycoside-Related Biosynthetic Enzymes, and K/Na Balance of Stevia under Salt Stress. Plants 2023, 12, 1628. [Google Scholar] [CrossRef]
  100. Ganjavi, A.S.; Oraei, M.; Gohari, G.; Akbari, A.; Faramarzi, A. Glycine betaine functionalized graphene oxide as a new engineering nanoparticle lessens salt stress impacts in sweet basil (Ocimum basilicum L.). Plant Physiol. Biochem. 2021, 162, 14–26. [Google Scholar] [CrossRef]
  101. Hanif, S.; Zia, M. Glycine betaine capped ZnO NPs eliminate oxidative stress to coriander plants grown under NaCl presence. Plant Physiol. Biochem. 2023, 197, 107651. [Google Scholar] [CrossRef] [PubMed]
  102. Dai, T.; Ban, S.; Han, L.; Li, L.; Zhang, Y.; Zhang, Y.; Zhu, W. Effects of exogenous glycine betaine on growth and development of tomato seedlings under cold stress. Front. Plant Sci. 2024, 15, 1332583. [Google Scholar] [CrossRef]
  103. Zhang, Y.; Dai, T.; Liu, Y.; Wang, J.; Wang, Q.; Zhu, W. Effect of Exogenous Glycine Betaine on the Germination of Tomato Seeds under Cold Stress. Int. J. Mol. Sci. 2022, 23, 10474. [Google Scholar] [CrossRef] [PubMed]
  104. Zhang, L.; Gao, M.; Hu, J.; Zhang, X.; Wang, K.; Ashraf, M. Modulation Role of Abscisic Acid (ABA) on Growth, Water Relations and Glycinebetaine Metabolism in Two Maize (Zea mays L.) Cultivars under Drought Stress. Int. J. Mol. Sci. 2012, 13, 3189–3202. [Google Scholar] [CrossRef] [PubMed]
  105. Sun, H.; Luo, M.; Zhou, X.; Zhou, Q.; Sun, Y.; Ge, W.; Wei, B.; Cheng, S.; Ji, S. Exogenous glycine betaine treatment alleviates low temperature-induced pericarp browning of ‘Nanguo’ pears by regulating antioxidant enzymes and proline metabolism. Food Chem. 2020, 306, 125626. [Google Scholar] [CrossRef]
  106. Zhao, J.; Lu, Z.; Wang, L.; Jin, B. Plant Responses to Heat Stress: Physiology, Transcription, Noncoding RNAs, and Epigenetics. Int. J. Mol. Sci. 2020, 22, 117. [Google Scholar] [CrossRef]
  107. Allakhverdiev, S.I.; Kreslavski, V.D.; Klimov, V.V.; Los, D.A.; Carpentier, R.; Mohanty, P. Heat stress: An overview of molecular responses in photosynthesis. Photosynth. Res. 2008, 98, 541–550. [Google Scholar] [CrossRef]
  108. Quan, J.; Li, X.; Li, Z.; Wu, M.; Zhu, B.; Hong, S.B.; Shi, J.; Zhu, Z.; Xu, L.; Zang, Y. Transcriptomic Analysis of Heat Stress Response in Brassica rapa L. ssp. pekinensis with Improved Thermotolerance through Exogenous Glycine Betaine. Int. J. Mol. Sci. 2023, 24, 6429. [Google Scholar] [CrossRef]
  109. Zhang, T.; Li, Z.; Li, D.; Li, C.; Wei, D.; Li, S.; Liu, Y.; Chen, T.H.H.; Yang, X. Comparative effects of glycinebetaine on the thermotolerance in codA- and BADH-transgenic tomato plants under high temperature stress. Plant Cell Rep. 2020, 39, 1525–1538. [Google Scholar] [CrossRef]
  110. Ali, S.; Abbas, Z.; Seleiman, M.F.; Rizwan, M.; YavaŞ, İ.; Alhammad, B.A.; Shami, A.; Hasanuzzaman, M.; Kalderis, D. Glycine Betaine Accumulation, Significance and Interests for Heavy Metal Tolerance in Plants. Plants 2020, 9, 896. [Google Scholar] [CrossRef]
  111. Khalid, M.; Rehman, H.M.; Ahmed, N.; Nawaz, S.; Saleem, F.; Ahmad, S.; Uzair, M.; Rana, I.A.; Atif, R.M.; Zaman, Q.U.; et al. Using Exogenous Melatonin, Glutathione, Proline, and Glycine Betaine Treatments to Combat Abiotic Stresses in Crops. Int. J. Mol. Sci. 2022, 23, 12913. [Google Scholar] [CrossRef] [PubMed]
  112. Sofy, A.R.; Dawoud, R.A.; Sofy, M.R.; Mohamed, H.I.; Hmed, A.A.; El-Dougdoug, N.K. Improving Regulation of Enzymatic and Non-Enzymatic Antioxidants and Stress-Related Gene Stimulation in Cucumber mosaic cucumovirus-Infected Cucumber Plants Treated with Glycine Betaine, Chitosan and Combination. Molecules 2020, 25, 2341. [Google Scholar] [CrossRef] [PubMed]
  113. Khanna, K.; Jamwal, V.L.; Sharma, A.; Gandhi, S.G.; Ohri, P.; Bhardwaj, R.; Al-Huqail, A.A.; Siddiqui, M.H.; Marraiki, N.; Ahmad, P. Evaluation of the role of rhizobacteria in controlling root-knot nematode infection in Lycopersicon esculentum plants by modulation in the secondary metabolite profiles. AoB Plants 2019, 11, plz069. [Google Scholar] [CrossRef]
  114. Khanna, K.; Sharma, A.; Ohri, P.; Bhardwaj, R.; Abd_Allah, E.F.; Hashem, A.; Ahmad, P. Impact of Plant Growth Promoting Rhizobacteria in the Orchestration of Lycopersicon esculentum Mill. Resistance to Plant Parasitic Nematodes: A Metabolomic Approach to Evaluate Defense Responses Under Field Conditions. Biomolecules 2019, 9, 676. [Google Scholar] [CrossRef]
  115. Niu, X.; Hu, Y.; Wang, X.; Li, R.; Li, W.; Shao, Y. Glycine betaine and MiWRKY53 enhance antioxidant capacity and disease resistance against Colletotrichum gloeosporioides in mango fruit. Postharvest Biol. Technol. 2025, 224, 113464. [Google Scholar] [CrossRef]
  116. Shan, T.; Jin, P.; Zhang, Y.; Huang, Y.; Wang, X.; Zheng, Y. Exogenous glycine betaine treatment enhances chilling tolerance of peach fruit during cold storage. Postharvest Biol. Technol. 2016, 114, 104–110. [Google Scholar] [CrossRef]
  117. Dong, X.; Ma, X.; Zhao, Z.; Ma, M. Exogenous betaine enhances salt tolerance of Glycyrrhiza uralensis through multiple pathways. BMC Plant Biol. 2024, 24, 165. [Google Scholar] [CrossRef]
  118. Min, K.; Cho, Y.; Kim, E.; Lee, M.; Lee, S.-R. Exogenous Glycine Betaine Application Improves Freezing Tolerance of Cabbage (Brassica oleracea L.) Leaves. Plants 2021, 10, 2821. [Google Scholar] [CrossRef]
  119. Hussain, N.; Yasmeen, A.; Yousaf, M. Antioxidant status and their enhancements strategies for water stress tolerance in chickpea. Braz. J. Biol. 2021, 82, e237809. [Google Scholar] [CrossRef]
  120. Wang, G.-Y.; Ahmad, S.; Wang, B.-W.; Shi, L.-B.; Wang, Y.; Shi, C.-Q.; Zhou, X.-B. Exogenous Glycinebetaine Regulates the Contrasting Responses in Leaf Physiochemical Attributes and Growth of Maize under Drought and Flooding Stresses. Biology 2024, 13, 360. [Google Scholar] [CrossRef]
  121. Chungloo, D.; Tisarum, R.; Samphumphuang, T.; Pipatsitee, P.; Sotesaritkul, T.; Cha-um, S. Exogenous glycine betaine alleviates water-deficit stress in Indian pennywort (Centella asiatica) under greenhouse conditions. Protoplasma 2024, 261, 625–639. [Google Scholar] [CrossRef]
  122. Wang, Y.-M.; Wang, Y.-N.; Shi, G.-L.; Liu, R.; Yang, A.; Ge, S.; Meng, H. Effects of exogenous betaine on physiological responses of peach tree under water stress. Chin. J. Appl. Ecol. 2007, 18, 542–548. [Google Scholar]
  123. Shemi, R.; Wang, R.; Gheith, E.-S.M.; Hussain, H.A.; Cholidah, L.; Zhang, K.; Zhang, S.; Wang, L. Role of exogenous-applied salicylic acid, zinc and glycine betaine to improve drought-tolerance in wheat during reproductive growth stages. BMC Plant Biol. 2021, 21, 574. [Google Scholar] [CrossRef] [PubMed]
  124. Manaf, H.H. Beneficial effects of exogenous selenium, glycine betaine and seaweed extract on salt stressed cowpea plant. Ann. Agric. Sci. 2016, 61, 41–48. [Google Scholar] [CrossRef]
  125. Hu, L.; Hu, T.; Zhang, X.; Pang, H.; Fu, J. Exogenous glycine betaine ameliorates the adverse effect of salt stress on perennial ryegrass. J. Am. Soc. Hortic. Sci. 2012, 137, 38–46. [Google Scholar] [CrossRef]
  126. Dong, X.; Liu, Y.; Ma, X.; Wang, S.; Yang, H.; Gao, X.; Wang, G.; Wang, H. Disclosing the effect of exogenous betaine on growth of Suaeda salsa (L.) Pall in the Liaohe coastal wetland, North China. Mar. Pollut. Bull. 2024, 198, 115852. [Google Scholar] [CrossRef]
  127. Zhao, Y.; Aspinall, D.; Paleg, L. Protection of membrane integrity in Medicago sativa L. by glycinebetaine against the effects of freezing. J. Plant Physiol. 1992, 140, 541–543. [Google Scholar] [CrossRef]
  128. Tarabih, M.; El-Eryan, E. Glycine Betaine and Proline with Thinning Technique for Resistance Abiotic Stress of Cristalina Cactus Pear. Pak. J. Biol. Sci. 2020, 23, 68–80. [Google Scholar] [CrossRef]
  129. Zheng, Z.; Wang, T.; Liu, M.; Xu, X.; Wang, J.; Sun, G.; He, S.; Liao, L.; Xiong, B.; Wang, X.; et al. Effects of Exogenous Application of Glycine Betaine Treatment on ‘Huangguoggan’ Fruit during Postharvest Storage. Int. J. Mol. Sci. 2023, 24, 14316. [Google Scholar] [CrossRef]
  130. Chen, W.; Li, P.; Chen, T. Glycinebetaine increases chilling tolerance and reduces chilling-induced lipid peroxidation in Zea mays L. Plant Cell Environ. Plant Cell Environ. 2000, 23, 609–618. [Google Scholar] [CrossRef]
  131. Oukarroum, A.; El Madidi, S.; Strasser, R. Exogenous glycine betaine and proline play a protective role in heat-stressed barley leaves (Hordeum vulgare L.): A chlorophyll a fluorescence study. Plant Biosyst. 2012, 146, 1037–1043. [Google Scholar] [CrossRef]
  132. Sorwong, A.; Sakhonwasee, S. Foliar application of glycine betaine mitigates the effect of heat stress in three marigold (Tagetes erecta) cultivars. Horticult. J. 2015, 84, 161–171. [Google Scholar] [CrossRef]
  133. Wang, G.-P.; Wang, J.-Z.; Xue, X.-M.; Lu, C.; Chen, R.; Wang, L.-P.; Han, X.-P. Foliar spraying of glycine betaine lowers photosynthesis inhibition of Malus hupehensis leaves under drought and heat stress. Int. J. Agric. Biol. 2020, 23, 1121–1128. [Google Scholar]
  134. Mo, X.; Qian, J.; Liu, P.; Zeng, H.; Chen, G.; Wang, Y. Exogenous Betaine Enhances the Protrusion Vigor of Rice Seeds under Heat Stress by Regulating Plant Hormone Signal Transduction and Its Interaction Network. Antioxidants 2022, 11, 1792. [Google Scholar] [CrossRef]
  135. Rasheed, R.; Wahid, A.; Farooq, M.; Hussain, I.; Basra, S.M.A. Role of proline and glycinebetaine pretreatments in improving heat tolerance of sprouting sugarcane (Saccharum sp.) buds. Plant Growth Regul. 2011, 65, 35–45. [Google Scholar] [CrossRef]
  136. Li, S.; Li, F.; Wang, J.; Zhang, W.; Meng, Q.; Chen, T.H.; Murata, N.; Yang, X. Glycinebetaine enhances the tolerance of tomato plants to high temperature during germination of seeds and growth of seedlings. Plant Cell Environ. 2011, 34, 1931–1943. [Google Scholar] [CrossRef] [PubMed]
  137. Wang, Y.; Liu, S.; Zhang, H.; Zhao, Y.; Zhao, H.; Liu, H. Glycine betaine application in grain filling wheat plants alleviates heat and high light-induced photoinhibition by enhancing the psbA transcription and stomatal conductance. Acta Physiol. Plant 2014, 36, 2195–2202. [Google Scholar] [CrossRef]
  138. Ahmad, R.; Ali, S.; Abid, M.; Rizwan, M.; Ali, B.; Tanveer, A.; Ahmad, I.; Azam, M.; Ghani, M.A. Glycinebetaine alleviates the chromium toxicity in Brassica oleracea L. by suppressing oxidative stress and modulating the plant morphology and photosynthetic attributes. Environ. Sci. Pollut. Res. Int. 2020, 27, 1101–1111. [Google Scholar] [CrossRef]
  139. Jabeen, N.; Abbas, Z.; Iqbal, M.; Rizwan, M.; Jabbar, A.; Farid, M.; Ali, S.; Ibrahim, M.; Abbas, F. Glycinebetaine mediates chromium tolerance in mung bean through lowering of Cr uptake and improved antioxidant system. Arch. Agron. Soil Sci. 2016, 62, 648–662. [Google Scholar] [CrossRef]
  140. Kumar, P.; Tokas, J.; Singal, H. Amelioration of chromium VI toxicity in Sorghum (Sorghum bicolor L.) using glycine betaine. Sci. Rep. 2019, 9, 16020. [Google Scholar] [CrossRef]
  141. Ali, S.; Chaudhary, A.; Rizwan, M.; Anwar, H.T.; Adrees, M.; Farid, M.; Irshad, M.K.; Hayat, T.; Anjum, S.A. Alleviation of chromium toxicity by glycinebetaine is related to elevated antioxidant enzymes and suppressed chromium uptake and oxidative stress in wheat (Triticum aestivum L.). Environ. Sci. Pollut. Res. Int. 2015, 22, 10669–10678. [Google Scholar] [CrossRef]
  142. Farooq, M.A.; Ali, S.; Hameed, A.; Bharwana, S.A.; Rizwan, M.; Ishaque, W.; Farid, M.; Mahmood, K.; Iqbal, Z. Cadmium stress in cotton seedlings: Physiological, photosynthesis and oxidative damages alleviated by glycinebetaine. S. Afr. J. Bot. 2016, 104, 61–68. [Google Scholar] [CrossRef]
  143. He, X.; Richmond, M.E.; Williams, D.V.; Zheng, W.; Wu, F. Exogenous glycinebetaine reduces cadmium uptake and mitigates cadmium toxicity in two tobacco genotypes differing in cadmium tolerance. Int. J. Mol. Sci. 2019, 20, 1612. [Google Scholar] [CrossRef]
  144. Stepien, P. Effects of the exogenous glycinebetaine on photosynthetic apparatus in cucumber leaves challenging Al stress. In Proceedings of the 18th International Conference on Heavy Metals in the Environment, Ghent, Belgium, 12–15 September 2016. [Google Scholar] [CrossRef]
  145. Ji, Y.; Ren, Y.; Han, C.; Zhu, W.; Gu, J.; He, J. Application of exogenous glycinebetaine alleviates lead toxicity in pakchoi (Brassica chinensis L.) by promoting antioxidant enzymes and suppressing Pb accumulation. Environ. Sci. Pollut. R. 2022, 29, 25568–25580. [Google Scholar] [CrossRef] [PubMed]
  146. Zhang, H.; Zhang, K.; Liu, T.; Zhang, Y.; Tang, Z.; Dong, J.; Wang, F. The characterization and expression analysis under stress conditions of PCST1 in Arabidopsis. Plant Signal Behav. 2022, 17, 2134675. [Google Scholar] [CrossRef] [PubMed]
  147. Jia, Z.; Wang, Y.; Wang, L.; Zheng, Y.; Jin, P. Amino acid metabolomic analysis involved in flavor quality and cold tolerance in peach fruit treated with exogenous glycine betaine. Food Res. Int. 2022, 157, 111204. [Google Scholar] [CrossRef] [PubMed]
  148. Yu, Z.; Niu, L.; Cai, Q.; Wei, J.; Shang, L.; Yang, X.; Ma, R. Improved salt-tolerance of transgenic soybean by stable over-expression of AhBADH gene from Atriplex hortensis. Plant Cell Rep. 2023, 42, 1291–1310. [Google Scholar] [CrossRef]
  149. Shahbazi, M.; Tohidfar, M.; Aliniaeifard, S.; Yazdanpanah, F.; Bosacchi, M. Transgenic tobacco co-expressing flavodoxin and betaine aldehyde dehydrogenase confers cadmium tolerance through boosting antioxidant capacity. Protoplasma 2022, 259, 965–979. [Google Scholar] [CrossRef]
  150. Imran, M.; Shafiq, S.; Tang, X. CRISPR-Cas9-mediated editing of BADH2 gene triggered fragrance revolution in rice. Physiol. Plant 2023, 175, e13871. [Google Scholar] [CrossRef]
  151. Phitaktansakul, R.; Kim, K.W.; Aung, K.M.; Maung, T.Z.; Min, M.H.; Somsri, A.; Lee, W.; Lee, S.B.; Nam, J.; Kim, S.H.; et al. Multi-omics analysis reveals the genetic basis of rice fragrance mediated by betaine aldehyde dehydrogenase 2. J. Adv. Res. 2022, 42, 303–314. [Google Scholar] [CrossRef]
  152. Hui, S.; Li, H.; Mawia, A.M.; Zhou, L.; Cai, J.; Ahmad, S.; Lai, C.; Wang, J.; Jiao, G.; Xie, L.; et al. Production of aromatic three-line hybrid rice using novel alleles of BADH2. Plant Biotechnol J. 2022, 20, 59–74. [Google Scholar] [CrossRef] [PubMed] [PubMed Central]
  153. Fitzgerald, T.L.; Waters, D.L.E.; Henry, R.J. The effect of salt on betaine aldehyde dehydrogenase transcript levels and 2-acetyl-1-pyrroline concentration in fragrant and non-fragrant rice (Oryza sativa). Plant Sci. 2008, 175, 539–546. [Google Scholar] [CrossRef]
  154. Prodhan, Z.H.; Islam, S.A.; Alam, M.S.; Li, S.; Jiang, M.; Tan, Y.; Shu, Q. Impact of OsBadh2 Mutations on Salt Stress Response in Rice. Plants 2022, 11, 2829. [Google Scholar] [CrossRef] [PubMed]
  155. Shi, J.-P.; Xie, R.-H.; Yang, F.; Zhou, W.; Jiang, X.-Q.; Hu, H.-F.; Yang, C.; Xie, Y.-Y.; Xiao, S. When agriculture meets biotechnology: A route for future agricultural innovation. Adv. Biotechnol. 2024, 2, 40. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Comparative schematic of the choline oxidation and glycine methylation pathways contributing to GB biosynthesis in plants and microbes. (A) The choline oxidation pathway in higher plants. (B) The glycine methylation pathway in microbes. (C) The serine metabolism pathway under stress. Abbreviations: 3-PGA, 3-phosphoglycerate; BADH, betaine aldehyde dehydrogenase; Bet Ald, betaine aldehyde; Cho, choline; CMO, choline monooxygenase; DMG, dimethylglycine; EA, ethanolamine; EK, ethanolamine kinase; GB, glycine betaine; Gly, glycine; GSMT, glycine/sarcosine N-methyltransferase; HPP, hydroxypyruvate; P-Cho, phosphocholine; P-DME, phosphodimethylethanolamine; P-EA, phosphoethanolamine; PEAMT, phosphoethanolamine N-methyltransferase; P-MME, phosphomonomethylethanolamine; Ptd-DME, phosphatidyl-dimethylethanolamine; Ptd-Cho, phosphatidylcholine; Ptd-EA, phosphatidylethanolamine; Ptd-MME, phosphatidyl-monomethylethanolamine; SAH, S-adenosylhomocysteine; SAM, S-adenosylmethionine; Sar, sarcosine; SDC, serine decarboxylase; SDMT, sarcosine/dimethylglycine N-methyltransferase; SHMT, serine hydroxymethyltransferase.
Figure 1. Comparative schematic of the choline oxidation and glycine methylation pathways contributing to GB biosynthesis in plants and microbes. (A) The choline oxidation pathway in higher plants. (B) The glycine methylation pathway in microbes. (C) The serine metabolism pathway under stress. Abbreviations: 3-PGA, 3-phosphoglycerate; BADH, betaine aldehyde dehydrogenase; Bet Ald, betaine aldehyde; Cho, choline; CMO, choline monooxygenase; DMG, dimethylglycine; EA, ethanolamine; EK, ethanolamine kinase; GB, glycine betaine; Gly, glycine; GSMT, glycine/sarcosine N-methyltransferase; HPP, hydroxypyruvate; P-Cho, phosphocholine; P-DME, phosphodimethylethanolamine; P-EA, phosphoethanolamine; PEAMT, phosphoethanolamine N-methyltransferase; P-MME, phosphomonomethylethanolamine; Ptd-DME, phosphatidyl-dimethylethanolamine; Ptd-Cho, phosphatidylcholine; Ptd-EA, phosphatidylethanolamine; Ptd-MME, phosphatidyl-monomethylethanolamine; SAH, S-adenosylhomocysteine; SAM, S-adenosylmethionine; Sar, sarcosine; SDC, serine decarboxylase; SDMT, sarcosine/dimethylglycine N-methyltransferase; SHMT, serine hydroxymethyltransferase.
Ijms 26 07971 g001
Figure 2. Mechanisms of GB action in regulating plant responses to multiple stresses (excessive light, pest infestations, temperature extremes, soil-related constraints). GB enhances plant tolerance to multiple stresses including drought, salinity, high temperatures, and insect herbivory. Abbreviation: ABA, abscisic acid; AOE, antioxidant enzymes; AsA, ascorbic acid; GA, gibberellic acid; GSH, glutathione; HKT, Na+/H+ antiporter; HMs, heavy metals; Pro, proline; ROS, reactive oxygen species; SA, salicylic acid; SS, soluble sugar.
Figure 2. Mechanisms of GB action in regulating plant responses to multiple stresses (excessive light, pest infestations, temperature extremes, soil-related constraints). GB enhances plant tolerance to multiple stresses including drought, salinity, high temperatures, and insect herbivory. Abbreviation: ABA, abscisic acid; AOE, antioxidant enzymes; AsA, ascorbic acid; GA, gibberellic acid; GSH, glutathione; HKT, Na+/H+ antiporter; HMs, heavy metals; Pro, proline; ROS, reactive oxygen species; SA, salicylic acid; SS, soluble sugar.
Ijms 26 07971 g002
Table 1. Summary of the application of glycine betaine on various plant species under abiotic stress.
Table 1. Summary of the application of glycine betaine on various plant species under abiotic stress.
Stress TypeSpeciesExogenous GB
Concentration
Resistance EffectReference
Drought stressChickpea (Cicer arietinum L.)0.17 mMIncreased antioxidant enzyme activity.[119]
Drought stressFlax (Linum usitatissimum)50–100 mMEnhancement on induction of protein, carbohydrate, and ROS scavenging systems.[91]
Drought stressMaize (Zea mays L.)0.5–10 mMImproves water retention and reduces osmotic potential, mitigating drought-induced water loss.[120]
Drought stressIndian pennywort (Centella asiatica)25–50 mMOverall physiological, morphological, and secondary metabolite traits were enhanced.[121]
Drought stressPeach (potted Prunus persica L.)0.85–4 mMDecreased leaf plasma membrane permeability, H2O2, Pro, and soluble sugar contents, increased leaf ASA-POD activity and soluble protein content.[122]
Drought stressSweet potato (Ipomoea batatas)50–100 mMIncreased sugar content and photosynthetic ability, controlled cellular osmotic potential, and maintained storage root yield.[70]
Drought stressTobacco (Nicotiana tabacum)20 mMImproved plant growth, osmotic adjustment, photosynthesis, and antioxidant enzyme activities.[56]
Drought stressWheat (Triticum aestivum)98.2 mMDecreased contents of H2O2, O2− and MDA, increased antioxidant enzymes, proline, and soluble sugar contents.[123]
Salt stressCowpeas (Vigna unguiculata)5–10 mMIncreased soluble sugar contents and antioxidant enzymes.[124]
Salt stressChinese licorice (Glycyrrhiza uralensis Fisch.)10–40 mMEnhances antioxidant defense, osmoregulation, and salt excretion capacity.[117]
Salt stressMaize (Zea mays L.)0.1 mMImproved photosynthesis and antioxidant activity.[97]
Salt stressPerennial ryegrass (Lolium perenne)20–50 mMIncreased fresh weight and relative water content; reduced electrolyte leakage and malondialdehyde content.[125]
Salt stressSuaeda salsa (L.)10–50 mMIncreased soluble sugars and elevated activity of Na+, K+-ATPase (enhancing osmotic stability).[126]
Cold stressAlfalfa (Medicago sativa)200 mMDecreased ion leakage from shoot tissues.[127]
Cold stressCabbage (Brassica oleracea L.)30 mMFreezing tolerance is enhanced by reducing membrane leakage, MDA, and ROS without affecting leaf growth.[118]
Cold stressCristalina Cactus Pear (OPUNTIA ficus-indica (L.) Mill.)5 mMImproved fruit quality through enhanced size, composition, and nutritional content.[128]
Cold stress‘Huangguogan’(Citrus reticulata Blanco)10–20 mMAntioxidant system activation reduces ROS and lipid peroxidation.[129]
Cold stressMaize (Zea mays L.)2.5 mMPrevented chlorosis; reduced lipid peroxidation of membrane.[130]
Cold stressPeach (Prunus persica Batsch.)10 mMActivating arginine/GABA metabolism and inhibiting polyamine degradation.[116]
Cold stressPears (Pyrus communis L.)10 mMDecreased membrane lipid peroxidation, maintained membrane integrity, increased activities and expression of APX, CAT, SOD, and Pro content.[105]
Cold stressTomato (Lycopersicon esculentum)10 mMMinimizes cold-induced seed damage by modulating oxidants, metabolites, and hormones to promote germination.[103]
Heat stressBarley (Hordeum vulgare)10 mMEnhances PSII stability by increasing antenna connectivity, protecting the oxygen-evolving complex.[131]
Heat stressMrigold (Tagetes erecta)0.5–1 mMImproves heat tolerance by protecting the photosynthetic apparatus, increasing stomatal conductance, and enhancing ROS scavenging.[132]
Heat stressPingyi Tiancha (Malus hupehensis (Pamp.) Rehder)10 mMImproved water status and enhanced antioxidant enzyme activity may underlie GB-induced enhancement of photosynthesis under stress.[133]
Heat stressRice (Oryza sativa L.)10 mMBoosting antioxidant enzymes, lowering MDA and ROS, and improving osmoregulation to ease heat stress.[134]
Heat stressSugarcane (Saccharum sp.)20 mMIncreasing K+ and Ca2+ levels, supporting tissue differentiation and dry weight accumulation.[135]
Heat stressTomato (Lycopersicon esculentum)0.1–5 mMEnhanced expression of heat-shock genes and accumulation of HSPs.[136]
Heat stressWheat (Triticum aestivum)100 mMMaintenance of higher chlorophyll content, PSII photochemical activity, and net photosynthetic rate.[137]
Heavy metal stress (Cr)Cauliflower (Brassica oleracea L.)1 mMIncreased dry biomass and improved antioxidative enzyme activities.[138]
Heavy metal stress (Cr)Mung bean (Vigna radiata)50–100 mMImproved plant growth; alleviated chromium stress.[139]
Heavy metal stress (Cr)Sorghum (Sorghum bicolor L.)50–100 mMIncreases antioxidant enzyme activity due to a decrease in chromium uptake or reduction in EL.[128,140]
Heavy metal stress (Cr)Wheat (Triticum aestivum)100 mMImproved growth, chlorophyll contents, and biomass and protein levels under chromium stress.[141]
Heavy metal stress (Cd)Cotton (Gossypium hirsutum Linn.)1 mMCd toxicity is mitigated by enhanced antioxidant enzyme activity.[142]
Heavy metal stress (Cd)Tobacco (Nicotiana tabacum)0.5 mMReduced MDA content, induced stomatal closure, improved leaf/root ultrastructure, increased the chl content, Fv/Fm, SOD, POD, CAT, and APX activities.[143]
Heavy metal stress (Al)Cucumber (Cucumis sativus L.)100 mMAccumulating in chloroplasts, GB mitigates Al stress by protecting the photosynthetic apparatus, enhancing electron transport, gas exchange, and CO2 fixation.[144]
Heavy metal stress (Pb)Pakchoi (Brassica campestris L.)0.5–2 mMIncreased dry biomass, mineral nutrient, and pigment contents, antioxidative enzyme activities, and reduced Pb contents.[145]
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

Chen, J.; Zhang, J.; Liu, Y.; Zhang, K.; Zhu, F.; Xie, Y. Advances in the Biosynthetic Regulation and Functional Mechanisms of Glycine Betaine for Enhancing Plant Stress Resilience. Int. J. Mol. Sci. 2025, 26, 7971. https://doi.org/10.3390/ijms26167971

AMA Style

Chen J, Zhang J, Liu Y, Zhang K, Zhu F, Xie Y. Advances in the Biosynthetic Regulation and Functional Mechanisms of Glycine Betaine for Enhancing Plant Stress Resilience. International Journal of Molecular Sciences. 2025; 26(16):7971. https://doi.org/10.3390/ijms26167971

Chicago/Turabian Style

Chen, Jiaxu, Jing Zhang, Yihang Liu, Kailu Zhang, Fuyuan Zhu, and Yanjie Xie. 2025. "Advances in the Biosynthetic Regulation and Functional Mechanisms of Glycine Betaine for Enhancing Plant Stress Resilience" International Journal of Molecular Sciences 26, no. 16: 7971. https://doi.org/10.3390/ijms26167971

APA Style

Chen, J., Zhang, J., Liu, Y., Zhang, K., Zhu, F., & Xie, Y. (2025). Advances in the Biosynthetic Regulation and Functional Mechanisms of Glycine Betaine for Enhancing Plant Stress Resilience. International Journal of Molecular Sciences, 26(16), 7971. https://doi.org/10.3390/ijms26167971

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