Next Article in Journal
Tansy (Tanacetum vulgare L.)—A Wild-Growing Aromatic Medicinal Plant with a Variable Essential Oil Composition
Next Article in Special Issue
Leaf Pigments, Surface Wax and Spectral Vegetation Indices for Heat Stress Resistance in Pea
Previous Article in Journal
Efficient Irrigation Methods and Optimal Nitrogen Dose to Enhance Wheat Yield, Inputs Efficiency and Economic Benefits in the North China Plain
Previous Article in Special Issue
MicroRNA and cDNA-Microarray as Potential Targets against Abiotic Stress Response in Plants: Advances and Prospects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Glycine Betaine in the Thermotolerance of Plants

1
Department of Horticultural Sciences, Faculty of Agriculture and Environment, The Islamia University of Bahawalpur, Bahawalpur 63100, Pakistan
2
Institute of Molecular Biology and Biotechnology, The University of Lahore, Lahore 54000, Pakistan
3
The UWA Institute of Agriculture, The University of Western Australia, Perth, WA 6001, Australia
*
Authors to whom correspondence should be addressed.
Agronomy 2022, 12(2), 276; https://doi.org/10.3390/agronomy12020276
Submission received: 18 November 2021 / Revised: 19 January 2022 / Accepted: 20 January 2022 / Published: 21 January 2022

Abstract

:
As global warming progresses, agriculture will likely be impacted enormously by the increasing heat stress (HS). Hence, future crops, especially in the southern Mediterranean regions, need thermotolerance to maintain global food security. In this regard, plant scientists are searching for solutions to tackle the yield-declining impacts of HS on crop plants. Glycine betaine (GB) has received considerable attention due to its multiple roles in imparting plant abiotic stress resistance, including to high temperature. Several studies have reported GB as a key osmoprotectant in mediating several plant responses to HS, including growth, protein modifications, photosynthesis, gene expression, and oxidative defense. GB accumulation in plants under HS differs; therefore, engineering genes for GB accumulation in non-accumulating plants is a key strategy for improving HS tolerance. Exogenous application of GB has shown promise for managing HS in plants, suggesting its involvement in protecting plant cells. Even though overexpressing GB in transgenics or exogenously applying it to plants induces tolerance to HS, this phenomenon needs to be unraveled under natural field conditions to design breeding programs and generate highly thermotolerant crops. This review summarizes the current knowledge on GB involvement in plant thermotolerance and discusses knowledge gaps and future research directions for enhancing thermotolerance in economically important crop plants.

1. Introduction

Temperature can adversely affect the normal functioning of plant metabolism [1,2]. In the last few decades, climate change-induced rising temperatures have beome a major challenge for modern crop production, especially in southern Mediterranean regions [3]. Thus, efforts to achieve maximum crop yields to ensure food security for an ever-increasing human population remain challenging. Global food security could be jeopardized if mitigation efforts are not implemented aggressively [4]. Any fluctuation in an environmental cue such as temperature can directly affect the production of temperature-sensitive crops, and hence food security, causing considerable losses to growers and other stakeholders. According to the Inter-Governmental Panel on Climate Change, global temperatures increased by 0.74 °C within a century (1906–2005), which was ascribed to ongoing anthropogenic activities and their resultant greenhouse gases [5]. The terrestrial surface temperature is expected to increase by a further 1–6 °C by 2050, with arable areas projected to see the greatest increases [6]. There is little doubt that the ongoing climate change will affect all societies inhabiting the globe [7,8], decreasing crop production, and threatening food security. Therefore, strategies are needed to ameliorate its effects on modern agriculture.
Heat stress is a condition of unfavorable temperature causing irreversible damage to plants [9]. High-temperature stress adversely affects plant physio-biochemical and molecular characteristics, resulting in poor plant growth and development [10]. At the physiological level, heat stress negatively influences photosynthesis by adversely affecting the oxygen evolving complex, photosystem II, RuBisCo, and energy-(ATP) producing processes [11,12]. Furthermore, heat stress-induced disturbance of the electron transport chain leads to excessive reactive oxygen species (ROS) production in different cell organelles, such as mitochondria and chloroplasts, causing severe damage to DNA and cell membranes by inducing lipid peroxidation and ultimately causing cell death 13]. Increasing the capacity of heat stress-induced excessive ROS scavenging is considered an efficient defense strategy for ameliorating heat stress in plants [13]. A plant’s thermotolerance ability is attributed to enhanced plasma membrane thermostability and reduced toxic ROS levels [14]. Plants have naturally adapted various defense mechanisms to counteract harsh environmental conditions such as heat stress. These defense mechanisms include an antioxidant machinery, osmolyte accumulation, maintenance of membrane integrity, and increased biosynthesis of heat-shock proteins (HSPs) by upregulating their associated genes’ expression [15,16]. These defense mechanisms are involved in cellular defense against heat stress. Osmolyte accumulation has a significant role in mediating stress tolerance in plants (reviewed in Zulfiqar et al. [17]). Various studies have reported enhanced GB accumulation in plants under heat stress, revealing the positive role of this osmolyte in heat stress tolerance [18,19,20,21,22]. Sorwong and Sakhonwasee [23] stated that exogenous GB application alleviated the heat stress-induced reduction in leaf gas exchange traits.
There are no critical reviews on the role of GB in heat stress tolerance. Here, we summarize the fundamental impact of GB in inducing heat stress tolerance in economically important crops.

2. GB Structure and Biosynthesis in Plants

Glycinebetaine is an N,N,N-trimethylglycine quaternary ammonium compound that is naturally produced in numerous living organisms, including plants [24]. At physiological pH, GB is electrically a neutral molecule, but dipolar in nature. There are two pathways related to GB biosynthesis in plants [25]. The initiating metabolites for these pathways are choline and glycine [26]. GB biosynthesis in plants occurs via N methylation of glycine or the oxidation of choline [27]. GB biosynthesis, particularly in higher plants, is a two-step pathway beginning with choline, which is catalyzed by a ferredoxin-dependent Rieske-type protein, namely, choline monooxygenase (CMO), and by a soluble NAD+-dependent enzyme [28,29]. Betaine aldehyde is oxidized by NAD+-dependent betaine aldehyde dehydrogenase (BADH) to produce GB. Both BADH and CMO generally exist in chloroplast stroma [24,30] (Figure 1).

3. Glycine Betaine-Accumulating and -Non-Accumulating Plants

Glycine betaine is a vital compatible osmolyte that plays multifarious roles in plant growth and metabolism. However, plant species differ in naturally accumulating GB. It is now evident that GB synthesis occurs in both chloroplasts and cytosol [32,33], but chloroplastic GB has been positively related to stress tolerance, whereas cytosolic GB has not shown such a relationship [32]. Thus, the presence of high amounts of GB in a plant may not necessarily account for its enhanced stress tolerance. Metabolic restriction of GB synthesis in plants has been ascribed to the availability of the substrate (choline) [33]. Since choline occurs in the cytosol [34], its transport to chloroplasts through appropriate transporters is essential for GB synthesis [34,35]. GB-non-accumulating plants have either limited amounts of intrinsic choline or poor activity of choline transporters in the chloroplast envelope [35]. Thus, genetic engineers need this information to generate transgenic lines with high GB-accumulating ability.
Plants accumulate various amounts of GB; naturally accumulating plants under normal and stress conditions are categorized as GB accumulators, while non-accumulating plants do not increase GB level under stress [36]. Table 1 lists GB accumulators and non-accumulators. Natural GB accumulators accrue a certain amount of GB solely under stressful cues [18,21]. For example, Alhaithloul et al. [22] studied the responses of Catharanthus roseus and Mentha piperita under HS and drought stress, individually and in combination. They reported that the level of GB increased by 46% and 58% for C. roseous and M. piperita, respectively, in response to HS, and more so under combined heat and drought stress. This evolutionary adaptation enables plants to survive under a varied range of climatic conditions. Screening plants for their ability to accumulate osmolytes, particularly GB, offers a way to target such plants for acquiring GB biosynthesis-related genes to develop GB-producing plants.

4. Mechanisms of GB-Mediated Thermotolerance

Glycine betaine is a compatible osmolyte that likely plays an important role in osmoregulation in plants subjected to extreme environmental cues, including high-temperature stress) [21,49]. Additionally, it is likely that GB activates signaling molecules such as calcium-dependent protein kinases (CDPKs) and mitogen-activated protein kinases (MAPKs) [50], which could activate stress-responsive and heat-shock transcription factor (HSF) genes [51,52]. The activated stress-responsive genes may boost the natural defense system by enhancing the activities of enzymatic antioxidants, such as superoxide dismutase (SOD), catalase (CAT), and peroxidase (POD), which may alleviate the negative impact of uncontrolled ROS causing oxidative damage triggered by heat stress)[53] (Figure 1). The elimination/reduction of ROS may keep biological membranes intact [54]. Furthermore, activated HSF genes may lead to the synthesis and activation of HSPs [55]. Most HSPs can also act as chaperones, which can prevent heat-induced aggregation of proteins [56]. The role of HSPs in plant thermotolerance has been elucidated in several comprehensive reviews [56,57]. GB can also significantly prevent photoinhibition by stabilizing the structure of the O2-evolving center (PSII) [19,58]. Thus, overall, GB can stabilize photosynthesis in heat-stressed plants, promoting growth under heat stress.
It is now evident that high temperatures cause many metabolic changes in plants that involve intricate reprogramming of cellular activities to safeguard organellar ultrastructures and functions under heat stress [59]. Although some promising roles of GB are depicted in Figure 1 for counteracting heat-induced physiological disorders, intensive research is needed to elucidate how and to what extent GB can regulate some key processes involved in plant thermotolerance, other than those highlighted in Figure 2. For example, very little information is available on the crosstalk between GB and other biomolecules, including various plant growth regulators.

5. Improvement in Heat Tolerance through Exogenous Application of GB

Exogenous application of GB improves thermotolerance in many plants by enhancing their growth and yield via counteracting metabolic maladjustments caused by HS (Table 2). For example, while appraising the role of exogenous GB application on heat-stressed tomato plants, Li et al. [60] reported enhanced seed germination, expression of heat-shock genes, and accumulation of heat-shock protein 70 (HSP70). Likewise, exogenous GB supplementation likely controls many other key metabolic processes in heat-stressed plants. For example, exogenously applied GB protected photosystem II (PSII) in heat-stressed plants of Hordeum vulgare [61] and Triticum aestivum [20] and decreased the relative membrane permeability and leakage of ions such as Ca2+, K+, and NO3 in heat-stressed barley seedlings [62]. Furthermore, GB supplied to sprouting sugarcane nodal buds under HS markedly reduced H2O2 generation, increased K+ and Ca2+ contents, and increased the levels of endogenous GB, free proline, and soluble sugars, enhancing the overall growth [63]. Sorwong and Sakhonwasee [23] stated that exogenous GB supplementation alleviated the heat stress-induced reduction in CO2 assimilation rate, stomatal conductance, relative water content, and transpiration rate in marigold. The high-temperature-induced oxidative stress in marigold was mitigated due to reduced levels of H2O2, peroxide, superoxide, and lipid peroxidation [23]. Exogenous application of GB during mid-flowering in heat-stressed tomato in the field increased fruit yield [64]. In apple, the application of GB enhanced photosynthesis under individual HS or drought stress and combined stresses [65]. In a three-year field study, Chowdhury et al. [66] evaluated the role of GB and potassium nitrate in heat-stressed late-sown wheat; these osmolytes improved grain yield under heat stress compared to the control. Hence, it is clear that exogenously applied GB mediates HS. However, future studies should focus on field-based heat stress evaluations of different crops.

6. Genetic Engineering for Enhanced Thermotolerance

Developing transgenic plants for thermotolerance is a cost-effective and efficient biotechnological approach for achieving optimum agricultural production under the changing climate scenario [17]. Genes involved in encoding GB biosynthetic enzymes in different organisms and plants have been cloned to produce transgenic plants overexpressing one or more of these genes to enhance endogenous GB production, improving HS tolerance [67,68] (Table 3). For example, Zhang et al. [68] compared the thermotolerance ability of two transgenic tomato lines containing the betaine aldehyde dehydrogenase (BADH) and choline oxidase (COD) genes responsible for GB synthesis. They observed that codA transgenic plants had higher GB levels, CO2 assimilation rate, and photosystem II (PSII) photochemical activity and lower accumulation of H2O2, O2•−, and malondialdehyde (MDA) than wild-type (WT) plants. In addition, the codA transgenic line had higher heat-response gene expression, heat-shock protein 70 (HSP70) accumulation, and expression of a mitochondrial small heat-shock protein (MT-sHSP), heat-shock cognate 70 (HSC70), and heat-shock protein 70 (HSP70) than WT plants under HS. In another study, Yang et al. [69] reported GB accumulation in tobacco by introducing the BADH gene in tobacco, which increased tolerance to high-temperature stress and improved photosynthesis. While transferring the BADH gene from spinach to tomato, Li et al. [67] reported enhanced accumulation of GB, antioxidant activity, and photosynthetic capacity by improving D1 protein content, which could repair heat stress-induced damage to PSII. Furthermore, transgenic tomato accumulated less MDA and ROS (H2O2 and O2•−), reducing oxidative stress relative to the WT. Reduced oxidative stress and restored PSII from HS-induced enhanced photoinhibition occurred in transgenic tobacco plants transformed with the BADH gene relative to the WT [70]. The role of BADH in xerophyllic Ammopiptanthus nanus under severe stress was confirmed by transferring the BADH gene of this plant into E. coli treated with 700 mM NaCl at 55 °C; the transgenic bacteria showed tolerance to these combined stresses [71]. The above studies reveal the positive role of GB-related genes in providing stress tolerance in plants. The introgression of GB synthesis-related genes could enhance endogenous GB accumulation to protect plants from heat-induced oxidative stress.
Numerous studies have been conducted on engineering GB biosynthesis [33]. Performance of single-gene-based transgenics under field conditions may not be the same as that reported under controlled or semi-controlled conditions. Thus, the development of transgenic lines by transforming with multiple genes (pyramiding of genes) for enhanced GB biosynthesis under heat stress is plausible.

7. Conclusions and Prospects

Under the changing climate threat, strategies are needed to alleviate the adverse impacts of harsh environmental stresses such as HS on plants. The most expedient strategy is to use the plant’s natural defense system for withstanding these stresses. Under HS, many plants naturally accumulate GB, a compound known to mediate HS tolerance via osmoregulation, photosynthetic mechanisms, and signaling molecules, such as CDPKs, MAPKs, nitric oxide (NO), and sugars, which activate stress-responsive genes and HSF genes.
As discussed above, GB biosynthesis has different roles in different organelles; for example, chloroplastic GB is actively involved in stress tolerance, while cytosolic GB lacks such functionality. As a result, high levels of GB in plants are not the only factor enhancing stress tolerance. The principal substrate for GB synthesis is choline, which occurs in the cytosol. Choline transport to the chloroplast takes place via choline transporters. Several problems occur in GB-non-accumulating plants: (1) a limited amount of endogenous choline exists and (2) choline transporters present on the chloroplast membrane do not transport the required level of choline to chloroplasts. Thus, molecular biologists require this information for different crops to develop transgenic lines/cultivars with enhanced GB-accumulating ability.
Plants that do not naturally accumulate GB under HS have less HS tolerance than those that do. Various strategies have been used to increase GB accumulation in these plants to improve their tolerance against HS. Exogenous supplementation of GB in different forms, such as seed priming or plant priming, has enhanced endogenous GB levels and thus improved plant growth and development under HS. Genetic engineering could be used to introduce biosynthetic pathway-related genes into GB-deficient species from plants or microorganisms. While various studies have demonstrated the importance of GB in thermotolerance, very few have revealed the molecular roles of GB in thermotolerance. Moreover, transgenic lines generated for different crops have been based on a single gene transformation, with marked progress in terms of enhanced GB accumulation coupled with improved thermotolerance. However, all these studies have been undertaken under semi-controlled or controlled conditions, and the developed transgenic lines have not been tested under natural field conditions. Thus, further research is needed to generate thermotolerant lines/cultivars for different crops threatened by the rising temperatures of climate change. The effectiveness of exogenous GB application should be tested at the field level.

Author Contributions

Conceptualization, F.Z. and M.A.; writing—original draft preparation; F.Z. writing—review and editing, F.Z., M.A. and K.H.M.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Geange, S.R.; Arnold, P.A.; Catling, A.A.; Coast, O.; Cook, A.M.; Gowland, K.M.; Leigh, A.; Notarnicola, R.F.; Posch, B.C.; Venn, S.E.; et al. The thermal tolerance of photosynthetic tissues: A global systematic review and agenda for future research. New Phytol. 2020, 229, 2497–2513. [Google Scholar] [CrossRef]
  2. Gil, K.E.; Park, C.M. Thermal adaptation and plasticity of the plant circadian clock. New Phytol. 2018, 221, 1215–1229. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wang, X.; Zhao, C.; Müller, C.; Wang, C.; Ciais, P.; Janssens, I.; Peñuelas, J.; Asseng, S.; Li, T.; Elliott, J.; et al. Emergent constraint on crop yield response to warmer temperature from field experiments. Nat. Sustain. 2020, 3, 908–916. [Google Scholar] [CrossRef]
  4. Sadok, W.; Lopez, J.R.; Smith, K.P. Transpiration increases under high-temperature stress: Potential mechanisms, trade-offs and prospects for crop resilience in a warming world. Plant Cell Environ. 2021, 44, 2102–2116. [Google Scholar] [CrossRef]
  5. IPCC. Climate change 2014: Impacts, adaptation and vulnerability. In Working Group II Contribution to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK, 2014. [Google Scholar]
  6. Ferguson, J.N.; Tidy, A.C.; Murchie, E.H.; Wilson, Z.A. The potential of resilient carbon dynamics for stabilizing crop reproductive development and productivity during heat stress. Plant Cell Environ. 2021, 44, 2066–2089. [Google Scholar] [CrossRef]
  7. O’Neill, B.C.; Carter, T.R.; Ebi, K.; Harrison, P.A. Achievements and needs for the climate change scenario framework. Nat. Clim. Chang. 2020, 10, 1074–1084. [Google Scholar] [CrossRef] [PubMed]
  8. Hertel, T.W.; de Lima, C.Z. Climate impacts on agriculture: Searching for keys under the streetlight. Food Policy 2020, 95, 101954. [Google Scholar] [CrossRef]
  9. Wahid, A.; Gelani, S.; Ashraf, M.; Foolad, M.R. Heat tolerance in plants: An overview. Environ. Exp. Bot. 2007, 61, 199–223. [Google Scholar] [CrossRef]
  10. Fahad, S.; Bajwa, A.A.; Nazir, U.; Anjum, S.A.; Farooq, A.; Zohaib, A.; Sadia, S.; Nasim, W.; Adkins, S.; Saud, S.; et al. Crop production under drought and heat stress: Plant responses and management options. Front Plant Sci. 2017, 8, 1147. [Google Scholar] [CrossRef] [Green Version]
  11. Tan, S.L.; Yang, Y.J.; Liu, T.; Zhang, S.B.; Huang, W. Responses of photosystem I compared with photosystem II to combination of heat stress and fluctuating light in tobacco leaves. Plant Sci. 2020, 292, 110371. [Google Scholar] [CrossRef]
  12. Parrotta, L.; Aloisi, I.; Faleri, C.; Romi, M.; Del Duca, S.; Cai, G. Chronic heat stress affects the photosynthetic apparatus of Solanum lycopersicum L. cv Micro-Tom. Plant Physiol. Biochem. 2020, 154, 463–475. [Google Scholar] [CrossRef] [PubMed]
  13. Suzuki, N.; Katano, K. Coordination between ROS regulatory systems and other pathways under heat stress and pathogen attack. Front Plant Sci. 2018, 9, 490. [Google Scholar] [CrossRef] [PubMed]
  14. Chen, W.L.; Yang, W.J.; Lo, H.F.; Yeh, D.M. Physiology, anatomy, and cell membrane thermostability selection of leafy radish (Raphanus sativus var. oleiformis Pers.) with different tolerance under heat stress. Sci. Hortic. 2014, 179, 367–375. [Google Scholar] [CrossRef]
  15. Akter, N.; Islam, M.R. Heat stress effects and management in wheat. A review. Agron. Sustain. Dev. 2017, 37, 37. [Google Scholar] [CrossRef]
  16. Waters, E.R.; Vierling, E. Plant small heat shock proteins–evolutionary and functional diversity. New Phytol. 2020, 227, 24–37. [Google Scholar] [CrossRef] [Green Version]
  17. Zulfiqar, F.; Akram, N.A.; Ashraf, M. Osmoprotection in plants under abiotic stresses: New insights into a classical phenomenon. Planta 2020, 251, 3. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Storey, R.; Ahmad, N.; Wyn Jones, R.G. Taxonomic and ecological aspects of the distribution of glycinebetaine and related compounds in plants. Oecologia 1977, 27, 319–332. [Google Scholar] [CrossRef]
  19. Allakhverdiev, S.I.; Los, D.A.; Mohanty, P.; Nishiyama, Y.; Murata, N. Glycinebetaine alleviates the inhibitory effect of moderate heat stress on the repair of photosystem II during photoinhibition. Biochim. Biophys. Acta (BBA)-Bioenerg. 2007, 1767, 1363–1371. [Google Scholar] [CrossRef] [Green Version]
  20. 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]
  21. 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] [Green Version]
  22. Alhaithloul, H.A.; Soliman, M.H.; Ameta, K.L.; El-Esawi, M.A.; Elkelish, A. Changes in ecophysiology, osmolytes, and secondary metabolites of the medicinal plants of Mentha piperita and Catharanthus roseus subjected to drought and heat stress. Biomolecules 2020, 10, 43. [Google Scholar] [CrossRef] [Green Version]
  23. Sorwong, A.; Sakhonwasee, S. Foliar application of glycine betaine mitigates the effect of heat stress in three marigold (Tagetes erecta) cultivars. Hort. J. 2015, 48, 161–171. [Google Scholar] [CrossRef] [Green Version]
  24. 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]
  25. Hanson, A.D.; Rhodes, D. 14C 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] [Green Version]
  26. Weretilnyk, E.A.; Bednarek, S.; McCue, K.F.; Rhodes, D.; Hanson, A.D. Comparative biochemical and immunological studies of the glycine betaine synthesis pathway in diverse families of dicotyledons. Planta 1989, 178, 342–352. [Google Scholar] [CrossRef]
  27. Chen, T.H.; Murata, N. Glycinebetaine: An effective protectant against abiotic stress in plants. Trends Plant Sci. 2008, 13, 499–505. [Google Scholar] [CrossRef] [PubMed]
  28. Brouquisse, R.; Weigel, P.; Rhodes, D.; Yocum, C.F.; Hanson, A.D. Evidence for a ferredoxin-dependent choline monooxygenase from spinach chloroplast stroma. Plant Physiol. 1989, 90, 322–329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  29. 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] [Green Version]
  30. Rathinasabapathi, B.; Burnet, M.; Russell, B.L. 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] [Green Version]
  31. Ashraf, M.; Foolad, M.R. Roles of glycine betaine and proline in improving plant abiotic stress resistance. Environ. Exp. Bot. 2007, 59, 206–216. [Google Scholar] [CrossRef]
  32. 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]
  33. Mansour, M.M.F.; Ali, E.F. Glycinebetaine in saline conditions: An assessment of the current state of knowledge. Acta Physiol. Plant. 2017, 39, 56. [Google Scholar] [CrossRef]
  34. 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]
  35. 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] [Green Version]
  36. Rhodes, D.; Hanson, A. Quaternary ammonium and tertiary sulfonium compounds in higher plants. Annu. Rev. Plant Physiol. Plant Mol. Biol. 1993, 44, 357–384. [Google Scholar] [CrossRef]
  37. Wyn Jones, R.G.; Storey, R. Betaines. In The Physiology and Biochemistry of Drought Resistance in Plants; Paleg, L.G., Aspinall, D., Eds.; Academic Press: Cambridge, MA, USA, 1981; pp. 171–204. [Google Scholar]
  38. Ladyman, J.A.R. The accumulation of glycinebetaine in barley in relation to water stress. Diss. Abs. Bull. 1982, 43, 1320. [Google Scholar]
  39. 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] [Green Version]
  40. Tian, F.; Wang, W.; Liang, C.; Wang, X.; Wang, G.; Wang, W. Overaccumulation of glycine betaine makes the function of the thylakoid membrane better in wheat under salt stress. Crop J. 2017, 5, 73–82. [Google Scholar] [CrossRef] [Green Version]
  41. Weimberg, R.; Lerner, H.R.; Poljakoff-Mayber, A. Changes in growth and water soluble solute concentrations in Sorghum bicolor stressed with sodium and potassium. Physiol. Plant. 1984, 62, 472–480. [Google Scholar] [CrossRef]
  42. Yang, W.-J.; Rich, P.J.; Axtell, J.D.; Wood, K.V.; Bonham, C.C.; Ejeta, G.; Mickelbart, M.V.; Rhodes, D. Genotypic variation for glycine betaine in sorghum. Crop Sci. 2003, 43, 162–169. [Google Scholar] [CrossRef]
  43. McCue, R.F.; Hanson, A.D. Drought and salt tolerance: Towards understanding and application. Tibtech 1990, 8, 358–362. [Google Scholar] [CrossRef]
  44. Kishitani, S.; Watanabe, K.; Yasuda, S.; Arakawa, K.; Takabe, T.J.P.C. Accumulation of glycinebetaine during cold acclimation and freezing tolerance in leaves of winter and spring barley plants. Plant Cell Environ. 1994, 17, 89–95. [Google Scholar] [CrossRef]
  45. 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]
  46. McNeil, S.D.; Nuccio, M.L.; Hanson, A.D. Betaines and related osmoprotectants. Targets for metabolic engineering of stress resistance. Plant Physiol. 1999, 120, 945–949. [Google Scholar] [CrossRef] [Green Version]
  47. 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] [PubMed] [Green Version]
  48. Mäkelä, P.; Munns, R.; Colmer, T.D.; Condon, A.G.; Peltonen-Sainio, P. Effect of foliar applications of glycinebetaine on stomatal conductance, abscisic acid and solute concentrations in leaves of salt-or drought-stressed tomato. Funct. Plant Biol. 1998, 25, 655–663. [Google Scholar] [CrossRef]
  49. Al-Huqail, A.; El-Dakak, R.M.; Sanad, M.N.; Badr, R.H.; Ibrahim, M.M.; Soliman, D.; Khan, F. Effects of climate temperature and water Stress on plant growth and accumulation of antioxidant compounds in sweet basil (Ocimum basilicum L.) leafy vegetable. Scientifica 2020, 2020, 3808909. [Google Scholar] [CrossRef] [Green Version]
  50. Hemantaranjan, A.; Bhanu, A.N.; Singh, M.N.; Yadav, D.K.; Patel, P.K.; Singh, R.; Katiyar, D. Heat stress responses and thermotolerance. Adv. Plants Agric. Res. 2014, 1, 62–70. [Google Scholar] [CrossRef] [Green Version]
  51. Divya, K.; Bhatnagar-Mathur, P.; Sharma, K.K.; Reddy, P.S. Heat shock proteins (Hsps) mediated signalling pathways during abiotic stress conditions. In Plant Signaling Molecules; Woodhead Publishing: Shaston, UK, 2019; pp. 499–516. [Google Scholar]
  52. Hashemi-Petroudi, S.H.; Nematzadeh, G.; Mohammadi, S.; Kuhlmann, M. Expression pattern analysis of heat shock transcription factors (HSFs) gene family in Aeluropus littoralis under salinity stress. Environ. Stresses Crop Sci. 2020, 13, 571–581. [Google Scholar]
  53. Zulfiqar, F.; Ashraf, M. Bioregulators: Unlocking their potential role in regulation of the plant oxidative defense system. Plant Mol. Biol. 2021, 105, 11–41. [Google Scholar] [CrossRef]
  54. Giri, J. Glycinebetaine and abiotic stress tolerance in plants. Plant Signal. Behav. 2011, 6, 1746–1751. [Google Scholar] [CrossRef]
  55. Andrási, N.; Pettkó-Szandtner, A.; Szabados, L. Diversity of plant heat shock factors: Regulation, interactions, and functions. J. Exp. Bot. 2020, 72, 1558–1575. [Google Scholar] [CrossRef] [PubMed]
  56. Jacob, P.; Hirt, H.; Bendahmane, A. The heat-shock protein/chaperone network and multiple stress resistance. Plant Biotechnol. J. 2017, 15, 405–414. [Google Scholar] [CrossRef]
  57. Reddy, P.S.; Chakradhar, T.; Reddy, R.A.; Nitnavare, R.B.; Mahanty, S.; Reddy, M.K. Role of heat shock proteins in improving heat stress tolerance in crop plants. In Heat Shock Proteins and Plants; Springer: Cham, Switzerland, 2016; pp. 283–307. [Google Scholar]
  58. Yang, G.; Rhodes, D.; Joly, R.J. Effects of high temperature on membrane stability and chlorophyll fluorescence in glycinebetaine-deficient and glycinebetaine-containing maize lines. Funct. Plant Biol. 1996, 23, 437–443. [Google Scholar] [CrossRef]
  59. Jagadish, S.K.; Way, D.A.; Sharkey, T.D. Plant heat stress: Concepts directing future research. Plant Cell Environ. 2021, 44, 1992–2005. [Google Scholar] [CrossRef] [PubMed]
  60. Li, S.; Li, F.; Wang, J.; Zhang, W.; Meng, Q.; Chen, T.H.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]
  61. Oukarroum, A.; El Madidi, S.; Strasser, R.J. Exogenous glycine betaine and proline play a protective role in heat-stressed barley leaves (Hordeum vulgare L.): A chlorophyll a fluorescence study. Plant Biosyst.—Int. J. Deal. All Asp. Plant Biol. 2012, 146, 1037–1043. [Google Scholar]
  62. Wahid, A.; Shabbir, A. Induction of heat stress tolerance in barley seedlings by pre-sowing seed treatment with glycinebetaine. Plant Growth Regul. 2005, 46, 133–141. [Google Scholar] [CrossRef]
  63. Rasheed, R.; Wahid, A.; Farooq, M.; Hussain, I.; Basra, S.M. 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]
  64. Makela, P.; Jokinen, K.; Kontturi, M.; Peltonen-Sainio, P.; Pehu, E.; Somersalo, S. Foliar application of glycine betaine—A novel product from sugar beet as an approach to increase tomato yield. Ind. Crops Prod. 1998, 7, 139–148. [Google Scholar] [CrossRef]
  65. Wang, G.; Wang, J.; Xue, X.; Lu, C.; Chen, R.; Wang, L.; Han, X. 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]
  66. Chowdhury, A.R.; Ghosh, M.; Lal, M.; Pal, A.; Hazra, K.K.; Acharya, S.; Chaurasiya, A.; Pathak, S.K. Foliar spray of synthetic osmolytes alleviates terminal heat stress in late-sown wheat. Int. J. Plant Prod. 2020, 14, 321–333. [Google Scholar] [CrossRef]
  67. Li, M.; Li, Z.; Li, S.; Guo, S.; Meng, Q.; Li, G.; Yang, X. Genetic Engineering of glycine betaine biosynthesis reduces heat-enhanced photoinhibition by enhancing antioxidative defense and alleviating lipid peroxidation in tomato. Plant Mol. Biol. Rep. 2014, 32, 42–51. [Google Scholar] [CrossRef]
  68. 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]
  69. Yang, X.H.; Liang, Z.; Lu, C.M. Genetic engineering of the biosynthesis of glycinebetaine enhances photosynthesis against high temperature stress in transgenic tobacco plants. Plant Physiol. 2005, 138, 299–309. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Yang, X.; Wen, X.; Gong, H.; Lu, Q.; Yang, Z.; Tang, Y.; Liang, Z.; Lu, C. Genetic engineering of the biosynthesis of glycinebetaine enhances thermotolerance of photosystem II in tobacco plants. Planta 2007, 225, 719–733. [Google Scholar] [CrossRef]
  71. Yu, H.Q.; Wang, Y.G.; Yong, T.M.; She, Y.H.; Fu, F.L.; Li, W.C. Heterologous expression of betaine aldehyde dehydrogenase gene from Ammopiptanthus nanus confers high salt and heat tolerance to Escherichia coli. Gene 2014, 549, 77–84. [Google Scholar] [CrossRef] [PubMed]
  72. Wang, G.P.; Li, F.; Zhang, J.; Zhao, M.R.; Hui, Z.; Wang, W. Overaccumulation of glycine betaine enhances tolerance of the photosynthetic apparatus to drought and heat stress in wheat. Photosynthetica 2010, 48, 30–41. [Google Scholar] [CrossRef]
Figure 1. Biosynthesis of GB in plant cells (Figure adapted from Sakamoto and Murata [24]; Ashraf and Foolad [31]).
Figure 1. Biosynthesis of GB in plant cells (Figure adapted from Sakamoto and Murata [24]; Ashraf and Foolad [31]).
Agronomy 12 00276 g001
Figure 2. Proposed mechanism of glycine betaine-mediated thermotolerance in plants.
Figure 2. Proposed mechanism of glycine betaine-mediated thermotolerance in plants.
Agronomy 12 00276 g002
Table 1. Glycine betaine accumulating crops.
Table 1. Glycine betaine accumulating crops.
GB AccumulatorsAccumulating ConditionReferences
Plant families with known naturally high accumulation of GB: Asteraceae, Chenopodiaceae, Poaceae, and SolanaceaeDifferent types of stresses[37,38,39]
Spinach (Spinacia oleracea)Naturally accumulates under non-stress conditions; GB levels increase under stress conditions[31,32,36]
Sugar beet (Beta vulgaris)Naturally accumulates under non-stress conditions; GB levels increase under stress conditions[31,32,36]
Barley (Hordeum vulgare)Naturally accumulates under non-stress conditions; GB levels increase under stress conditions[31,36,40]
Wheat (Triticum aestivum)Naturally accumulates under non-stress conditions; GB levels increase under stress conditions[31,36,40]
Sorghum (Sorghum bicolor)Naturally accumulates under non-stress conditions; GB levels increase under stress conditions[41,42]
Maize (Zea mays)Naturally accumulates under non-stress conditions; GB levels increase under stress conditons[36,43,44,45]
GB-non-accumulators
Rice (Oryza sativa)Non-stress and stress conditions[31,36,46,47]
Mustard (Brassica spp.)Non-stress and stress conditions[46]
Arabidopsis (Arabidopsis thaliana)Non-stress and stress conditions[33,36,46]
Tobacco (Nicotiana tabacum)Non-stress and stress conditions[33,46]
Tomato (Solanum lycopersicum)Non-stress and stress conditions[33,36,46,48]
Potato (Solanum tuberosum)Non-stress and stress conditions[36,46]
Table 2. Effect of exogenously applied GB on the regulation of different physio-biochemical attributes in heat-stressed plants.
Table 2. Effect of exogenously applied GB on the regulation of different physio-biochemical attributes in heat-stressed plants.
CropHeat Stress RangeGB Concentration AppliedExogenously Applied GB-Induced Regulation of Different Attributes in Heat-Stressed PlantsReference
Tomato (Lycopersicon esculentum)34 °C0, 0.1, 1, and 5 mM GB
  • Improved seed germination
  • Enhanced expression of heat-shock genes and accumulation of HSPs
[60]
Barley (Hordeum vulgare)45 °C10 mM
  • Protective effect on oxygen-evolving complex by increasing connectivity among PSII antennae, inducing greater PSII stability
[61]
Wheat (Triticum aestivum)25/20 °C day/night100 mM
  • Maintenance of higher chlorophyll content, PSII photochemical activity, and net photosynthetic rate
  • Reversed the decline in psbA gene transcription
  • Accelerated endogenous accumulation of GB in leaves
  • Decreased photoinhibition by D1 protein synthesis
[20]
Wheat (T. aestivum)30–38 °C100 and 50 mM
  • Improved yield
  • Marginally improved the relative membrane permeability
[66]
Barley (H. vulgare)40/32 °C day/night10, 20, 30, 40, and 50 mM
  • Improved growth, photosynthesis, and water relations
  • Decrease ion leakage
[62]
Sugarcane (Saccharum spp.)42 °C20 mM
  • Improved bud sprouting
  • Decreased H2O2 production
  • Improved soluble sugar accumulation
  • Improved K+ and Ca2+ content
  • Enhanced the endogenous levels of osmolytes
[63]
Marigold (Tagetes erecta)39/29 °C day/night0.5 and 1 mM
  • Improved leaf gas exchange traits
  • Decreased ROS accumulation
[23]
Abbreviations: HSPs: Heat-shock proteins, PSII: Photosystem II, ROS: Reactive oxygen species, psbA: PSII protein D1 precursor gene.
Table 3. Genetic engineering for enhanced GB accumulation and improved thermotolerance in different crops.
Table 3. Genetic engineering for enhanced GB accumulation and improved thermotolerance in different crops.
Gene TransformedDonor/SourceGene ActionTransgenic Plant SpeciesStress ConditionTransgenic Plant ResponseReferences
BADH and codASpinach (Spinacia oleracea L.) as a donor of the BADH gene; binary vector pCG/codA for chloroplast-targeted expression of the codA geneGenes related to key enzymes involved in GB synthesisTomato (Solanum lycopersicum)Two months after transplanting, plants were exposed to 42 °C for 0–8 h in a growth chamber
  • Enhanced GB levels in leaves
  • Improved CO2 assimilation and photosystem II (PSII) photochemical activity
  • Transgenic plants, especially those containing codA, accumulated low ROS levels and thus exhibited reduced oxidative stress
  • Enhanced expression of heat-response genes and accumulation of heat-shock protein 70 (HSP70)
  • Enhanced thermotolerance
[68]
BADHSpinach (S. oleracea L.)Gene for betaine aldehyde dehydrogenaseTobacco (Nicotiana tabacum)Two-month-old seedlings subjected to various temperatures (25–50 °C) for 4 h in a growth chamber
  • Increased GB accumulation in vivo increased PSII tolerance under heat stress
  • Enhanced GB accumulation in vivo improved the thermostability of PSII reaction centers
  • Reversed heat-induced PSII photoinhibition
  • Accumulated low levels of ROS, improving oxidative defense system
[70]
BADHSpinach (S. oleracea L.)Gene for betaine aldehyde dehydrogenaseTomato (S. lycopersicum)Six-week-old seedlings placed in a growth chamber at 42 °C for 0, 2, 4, or 6 h
  • Higher GB accumulation
  • Enhanced chlorophyll fluorescence
  • Increased tolerance to heat-induced photoinhibition
  • Improved D1 protein content
  • Enhanced antioxidant enzyme activities
  • Reduced oxidative stress
[67]
BADHSpinach (S. oleracea L.)---do---Tomato (Solanum lycopersicum)Two-month-old plants exposed to varying temperature regimes (25–45 °C) for 2 h
  • Enhanced CO2 assimilation
  • Enhanced photosynthesis related to RuBisCo activase-mediated activation of RuBisCo
  • Enhanced thermotolerance
[69]
BADHGarden orache (Atriplex hortensis L.)--do--Wheat (Triticum aestivum)Individual and combined heat stress (40 °C) and drought stress as PEG-6000 (osmotic potential about −1.88 MPa) for 3 h in an artificial chamber
  • Improved photosynthesis
  • Improved antioxidant activities/levels
  • Improved water status
[72]
codASpinach (S. oleracea L.)Gene encodescholine oxidase (COD), a key enzyme for GB synthesisRice (Oryza sativa)Plants grown at 28/13 °C for 5 weeks
  • Enhanced endogenous GB accumulation
  • Improved heat and salt stress tolerance
[47]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zulfiqar, F.; Ashraf, M.; Siddique, K.H.M. Role of Glycine Betaine in the Thermotolerance of Plants. Agronomy 2022, 12, 276. https://doi.org/10.3390/agronomy12020276

AMA Style

Zulfiqar F, Ashraf M, Siddique KHM. Role of Glycine Betaine in the Thermotolerance of Plants. Agronomy. 2022; 12(2):276. https://doi.org/10.3390/agronomy12020276

Chicago/Turabian Style

Zulfiqar, Faisal, Muhammad Ashraf, and Kadambot H. M. Siddique. 2022. "Role of Glycine Betaine in the Thermotolerance of Plants" Agronomy 12, no. 2: 276. https://doi.org/10.3390/agronomy12020276

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