Next Article in Journal
Improving Dengue Forecasts by Using Geospatial Big Data Analysis in Google Earth Engine and the Historical Dengue Information-Aided Long Short Term Memory Modeling
Previous Article in Journal
Retinoic Acid-Differentiated Neuroblastoma SH-SY5Y Is an Accessible In Vitro Model to Study Native Human Acid-Sensing Ion Channels 1a (ASIC1a)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Seed and the Metabolism Regulation

by
Hayat El-Maarouf-Bouteau
Institut de Biologie Paris Seine (IBPS), Sorbonne Université, UMR 7622, 4, Place Jussieu, 75005 Paris, France
Submission received: 13 December 2021 / Revised: 13 January 2022 / Accepted: 18 January 2022 / Published: 20 January 2022
(This article belongs to the Special Issue Current Advances in Seed Biology)

Abstract

:

Simple Summary

Seeds are the reproductive units of higher plants. They have a significant place in agriculture and plant diversity maintenance. Because they are dehydrated, they can remain viable in the environment for centuries. This review explores the dry seed as a metabolically inactive organism, but well organized to protect its components and enter intensive repair to restore metabolic activities upon imbibition for the completion of germination. Metabolism regulation is also critical for the most important seed traits, dormancy, and ageing recovery capacity.

Abstract

The seed represents a critical stage in the life cycle of flowering plants. It corresponds to a dry structure carrying the plant embryo in dormant or quiescent state. Orthodox seeds possess a very low water content, preventing biochemical reactions, especially respiration. If the desiccation of living organisms leads to a loss of homeostasis, structure, and metabolism, the seeds go through it successfully thanks to their structure, cellular organization, and growth regulation. Seeds set up a certain number of sophisticated molecules to protect valuable macromolecules or organelles from dehydration/rehydration cycles. Moreover, dormancy takes place in a coordinated process with environmental cues in order to ensure embryo development at the most appropriate conditions for the establishment of the new plant. Moreover, repair processes are programmed to be ready to operate to maximize germination success and seed longevity. This review focuses on the physiology of the seed as related to hydration forces, respiration, and biochemical reactions in the transition from thermodynamically undefined dry state to self-sustained living system. Such processes are of importance for basic knowledge of the regulation of metabolism of living organisms, but also for the control of germination in the context of climate change due to global warming.

1. Introduction

Seeds are important as propagation units for crops, but also for species maintenance in the natural environment. Seed germination represents the first step in the establishment of the new plant for agriculture or in natural areas. It is, therefore, important to unravel the physiological aspects of germination for basic knowledge, as well as for the good management in the context of environmental fluctuations due to global warming. Understanding germination depends on understanding the seed organization and functioning in anhydrobiosis [1]. In fact, the most important characteristic of the seeds referred to as orthodox, which are the focus of the present review, is the ability to be desiccated and to survive dry state, allowing them to be stored and distributed widely. On the contrary, recalcitrant seeds cannot tolerate dehydration. They possess a high water content and active metabolism and cannot be stored for long periods [2]. Another category with intermediate features also exists, e.g., coffee seeds, which can tolerate drying but display sensitivity to cool temperatures [3].
In the absence of metabolic activity, orthodox seeds do not meet the different definitions of living organisms. The biological definition of living organisms admitted so far is based on the ability of regeneration and the existence of metabolism. NASA’s defition of life based on thermodynamic law is a “self-sustaining chemical system capable of Darwinian evolution” [4]. In both cases, living organisms can be characterized by metabolic activity having an interaction with ecological conditions. The self-sustaining chemical system corresponds to a thermodynamic aspect of life as a system far from equilibrium [5]. Yet, the seed carries the embryo ready to live as soon as the seed rehydrates. In this review, seed organization as related to water status and seed metabolism in dry state and upon imbibition are examined to understand dormancy, germination, and ageing tolerance processes.

2. Dry Seed: Well-Organized to Resist

2.1. The Seed, a Special New Individual

The seed is composed of an embryo surrounded by reserve material and covering layers. It represents the plant dispersion organ formed by sexual reproduction as well as the new individual. The seed therefore occupies a critical position in the life cycle of the higher plant. The success of the establishment of the new individual is determined by physiological and biochemical features of the seeds in response to their environment.
The angiosperm seed generally consists of the embryo, the result of the fertilization of the egg cell and one of the male pollen nuclei, the endosperm, which is the result of the fusion of the two polar nuclei with the second pollen nuclei, and the perisperm, corresponding to the nucellus and the testa or seed coat formed from the integument around the ovule. The extent to which the endosperm or perisperm persists varies between species. For example, the Arabidopsis embryo is surrounded by an endosperm layer while the sunflower embryo is not (Figure 1). When the testa is underdeveloped, the outer structure being the pericarp or fruit coat, the dispersal unit is not a seed but a fruit, as in the case of sunflower and wheat. The embryo, which represents the new individual, is comprised of the embryonic axis and one or two cotyledons. The axis includes the embryonic root (radicle), the hypocotyl, and the shoot apex (plumule). Thus, as the seed corresponds to a diverse composition of such complex tissues that have distinct developmental programs [6], studying seeds implies the study of these programs and their coordination in time and space to achieve germination [7].
Desiccation represents the last phase of seed development and corresponds to a huge loss of water content that decreases the seed water percentage in orthodox seeds to less than 10% of the dry weight (DW), depending on species (e.g., mature sunflower seeds contain 4% g H2O/g DW [1]). Such a low water content changes the cytoplasm from a fluid to glassy state, which severely reduces molecular diffusion and mobility, preventing chemical reactions [8]. In fact, at dry state, cellular metabolism and respiration are greatly reduced [9,10]. Thus, dry seeds maintain low levels of metabolic activity, which preserves their viability for years or even centuries, as for Phoenix dactylifera L. seeds [11]. The mechanisms by which the seed tolerates desiccation are discussed in specialized reviews [12,13]. In this review, the focus will be on physiological changes allowing mature seeds to successfully undertake conservation and germination.
Seed germination starts with water uptake and ends with radicle protrusion. The seed water absorption rate corresponds to three phases during which controlled physiological processes take place. As shown in the Figure 2, phase I corresponds to a rapid water uptake, which induces the transformation of cell membranes from gel phase to liquid crystal state and the reorganization of cell structure and molecules required for the establishment of cell metabolism that takes place actively at constant water content corresponding to phase II (plateau phase), during which the water uptake is stopped. In fact, based on reports on different species, such as wheat, rice, Arabidopsis, and sunflower, phase II corresponds to high metabolic activity, with gene expression corresponding to respiration, hormones, sugar, and cell wall metabolism, and protein turn-over allowing repair and component preparation for cell elongation and growth [14,15,16,17,18]. During phase III, fast water uptake takes place again to ensure reserve mobilization and metabolism for root elongation and growth [19].
However, germination is not generally possible for mature seeds because they are generally dormant. They need to undergo a post-maturation phase, called after-ripening, a period that allows them to acquire the capacity to germinate. The transition of dry seeds from dormant (D) to non-dormant (ND) state corresponds to determinant physiological changes from arrested to permissive processes leading to germination. The characterization of possible chemical reactions and subsequent physiological activity at dry state remain the most difficult question in seed biology because experimental procedures require short- or long-term hydration. Yet, this question is crucial in the understanding of dormancy alleviation, germination, and longevity.

2.2. Water, “Matrix of Life”

If water is the matrix of life [20], dry seeds can hardly be considered as alive and yet they bear life in the form of the embryo. Water is an essential participant in the chemistry of life by sustaining the biochemistry of the cell. It acts as a liquid and solvent for biochemical reactions, but also influences macromolecule structures [21]. Water participates in the catalytic function of proteins and nucleic acids and physically in hydrophobic associated protein folding and complex formation through the hydrogen bond [22]. Physical methods, such as thermodynamic studies or nuclear magnetic resonance spectroscopy, came to the rescue of biology for the investigation of water status in low hydrated seeds and subsequent interactions. Using thermodynamic measurement, three levels of water affinity have been characterized in pea and soybean seeds [9,23,24]. Strongly bound water was recorded at 8% of water content, weakly bound water between 8% and 24%, and very loosely bound water at contents above 24% [9]. In these ranges of water moisture, the investigation of lysozyme hydration by IR spectroscopy and heat capacity showed that with up to 0.07 g of water/g of protein, the hydration process is dominated by the interaction with charged groups. At 0.07 g/g, there is a transition in the IR spectrum and the heat capacity, reflecting a change in surface water arrangements. Between 0.07 and 0.25 g/g, most of the surface is covered with water molecules. Between 0.25 and 0.38 g/g, water condenses over the non-polar atoms not adjacent to charged or polar atoms [25]. The final stage of protein hydration is that of hydrophobic groups, which represent a large portion of the surface of the protein molecule. Water–water bonds can be then created and participate in protein–protein or protein–substrate interactions. The enzymatic activity of lysozymes becomes detectable at 0.2 g/g and changes with hydration above 0.38 g/g. Changes in the arrangement of water in the protein environment affect protein stability and enzyme properties [25,26]. On the other hand, nucleic acids require more water than proteins [25]. In fact, the end point of the hydration process of nucleic acids is about twice the level for proteins [27]. DNA structure and related biological functions are controlled by the complex dynamics of hydrating water and ions in and around the DNA [28]. It was shown that in desiccated Arabidopsis seeds the chromatin is highly condensed and can be de-condensated after hydration [29]. The property of the seed to undergo a reversible chromatin condensation/de-condensation enables to withstand desiccation and the entry in active metabolism during imbibition.
Thus, high water binding forces in dry seeds are responsible for the lack of stability and activity of biomolecules causing low metabolism and energy. As membrane reorganization is one of the first events in the initiation of cell energization, more water (>24%) is needed to activate protein reorganization and activity for full plasma membrane and mitochondrial energy restoration (Figure 3). Indeed, below 24% of water, seed O2 consumption is very low and it is undetectable at around 8% [9]. Respiration plays a crucial role in providing cellular energy via oxidative phosphorylation, but it also represents the major source of reactive oxygen species (ROS) responsible for cell damage. By preventing cell metabolism, drying keeps the embryo alive, which highlights the dual role of water in life and its consumption.

2.3. Respiration Resumption

Holding respiration may represent the major process allowing seed longevity. This is achieved with the support of several seed features. The seed structure itself may contribute to holding respiration by O2 uptake limitation due to the space occupation by reserve molecules and the presence of the seed coat (for review [13,30]). More importantly, mitochondria in dry seeds are called promitochondria as their internal membranes are underdeveloped with a low number of cristae and low protein content [31,32]. Several studies using different biochemical approaches, such as adenylate pool or adenylate energy charge (AEC) ratio ((ATP + 0.5 ADP)/(AMP + ADP + ATP)), oxygen uptake, tricarboxylic cycle enzyme activity measurements, or cytological investigations, converge to state that respiration is reduced to a very low level in dry seeds and that the hydration induces an increase in mitochondria components and activity [33,34,35,36,37,38,39]. The proliferation and differentiation of mitochondria, called ‘mitochondrial biogenesis’, occur progressively upon imbibition [10,37,38]. These are considered as the prerequisite for the full reactivation of mitochondria and subsequent energy supply for germination [31,32]. However, isolated promitochondria were shown to be able to generate ATP and a membrane potential by oxidizing supplied succinate and/or NADH [31,40]. Such metabolic activity may be decisive at the onset of imbibition to help the biogenesis process to take place. Moreover, promitochondria seem to have an import apparatus ready for mitochondrial biogenesis [32]. In fact, the electron transport system is activated immediately after the initiation of imbibition and is dependent on AMP, ADP, cytochrome C oxidase, and ATPase that were recovered from dry seeds [36,41]. A recent study enabled the visualization of mitochondrial reactivation and the chondriome (all mitochondria in a cell) during imbibition [10]. They confirmed that promitochondria have reduced metabolic activity but can generate a membrane potential within the first minutes of imbibition. Further imbibition in permissive conditions for germination allowed a significant increase of mitochondrial dynamics, leading to inter-mitochondria interactions and localization around the nucleus, which may facilitate mitochondrial biogenesis and synchronization [10].
Mitochondrial functioning is also dependent on post-translational modifications of proteins of metabolic functions. In fact, the NADPH produced from the metabolism enables the reduction of thiol redox reactions [42]. Thus, mitochondrial resumption enables not only the production of energy as ATP for elongation and growth, but also reductants that determine redox regulation for subsequent transcription and hormonal regulation.

2.4. Plasma Membrane Potential

One of the fundamental properties of living cells is the establishment of an electrical potential difference across the plasma membrane. In dry seeds, the transport of ions is not possible due to the absence of water as a conductive fluid, but also due to the loss of the integrity of membranes and their protein components. Membrane deterioration has been highlighted by the high electrolyte leakage rate measured in dry seeds and many studies have reported that seed hydration induced a membrane leakage decrease, confirming that the cell membrane was repaired upon imbibition. It was shown that the leakage decrease depends on the moisture content of the seeds, being undetectable above 24% [9]. In fact, 20% water corresponds to the minimum amount of water needed to create a hydrophilic layer that stabilizes the organization of lipids in a bilayer [43]. Such a water content corresponds to phase I of germination (Figure 3), while lipid metabolism and repair were shown to take place during phase II [44,45], suggesting the need for an intact membrane before any repair event can be started. It also corresponds to the start point of the full hydration of proteins (discussed above) that should include membrane ion channels and transporters.
At a water content less than 20%, the cell membrane may consist of fragments of a hexagonal array of hydrophilic circles formed by polar heads of phospholipids [43]. Such organization is responsible for electrolyte leakage and probably facilitates the massive entry of water at the beginning of imbibition. The characterization of these electrolytes in several seeds showed a great diversity of molecules, such as ions, amino-acid, sugars, organic acids, phenols, and phosphates, as well as hormones like gibberellic acid [43]. If the membrane disorganization of the dry seed inevitably induces the release of electrolytes at the beginning of imbibition, it would correspond to a powerful process which allows the seed to germinate on the poorest supports by modifying the external environment charges to create a membrane electrical potential.
In plants, plasma membrane potential is driven by two major components, K+ gradient and H+ ATPase activity. The plasma membrane (PM) H+ ATPase is responsible for membrane energization by extruding H+ protons, which is necessary for the activity of nutrient transporters associated to electrochemical H+ gradient [46]. It was demonstrated that PM H+ ATPase is essential for growth since the knockout of the two major PM H+-ATPase genes, AHA1 and AHA2, is lethal in Arabidopsis embryos [47]. The role of PM H+ ATPase in physiological processes is regulated by post-translational modifications which correspond to the phosphorylation of C terminus residues [48,49]. It was shown that PM H+-ATPase presents two activity states, auto-inhibited and upregulated, depending on the coupling ratio between ATP hydrolysis and H+ pumping [48,50,51]. The basal state has a low coupling ratio, while the activated state has a high ratio [51]. Several signals, such as sugar or light, activate the phosphorylation of C terminus, allowing the activation state corresponding to high affinity for ATP [48]. Although H+-ATPase has not been actively studied in seeds, recent work has shown that high H+-ATPase activity was associated with germination capacity while dormant state was associated with low activity in sunflower seeds [52]. Considering that the imbibition of dry seeds is driven by the physical properties of water in the reorganization and remodeling of PM, including the proper folding of H+-ATPase as a protein component, and given the central role of mitochondria and reserve mobilization, the ATP/ADP ratio of the cell may be the major parameter affecting PM H+-ATPase activity in the seed germination process. Further investigations are needed to discover the pathways by which this protein is phosphorylated and dephosphorylated in the regulation of dormancy and germination.

3. Seed Dormancy: Higher Level of Resistance

Seed dormancy, which is the incapacity of mature seeds to germinate, is one of the most important processes in the successful establishment of the new seedling. Dormancy is finely regulated with the aim to insure germination at the optimal moment. Indeed, deep dormancy prevents field emergence and low dormancy causes sprouting. Dormancy takes place at the end of seed formation, and it’s removed during a period of several weeks to decades, called after-ripening.

3.1. Seed Metabolism and Dormancy

After-ripening has fascinated researchers because dormancy is alleviated in dry conditions, suggesting that some processes operate in the dry seed. Biological reactions have been investigated and transcriptional programs have been proposed to be involved in the regulation of after-ripening-mediated seed dormancy alleviation in several seeds [53,54,55,56,57]. Given the restricted molecular mobility due to the glassy state in dry seed cells, the existence of a hydrated pocket within the cell enabling gene transcription has been hypothesized [53]. To address this issue, Meimoun et al. [1] investigated transcriptomic changes after the after-ripening period in sunflower seeds using two protocols, one allowing dormancy alleviation but not the other, in order to differentiate between changes in gene expression associated with dormancy alleviation and those associated with storage only. They showed that there is no significant variation between conditions, suggesting that gene expression did not take place during after-ripening, in agreement with the absence of metabolic activity in dry seeds [1]. Furthermore, ancient studies (over 50 years) have already shown that transcription was not required for de novo protein synthesis in imbibed seeds, suggesting that seeds contain stored transcripts ready for use upon imbibition [58]. Since then, a number of studies have demonstrated that germination (reaching radicle protrusion) is completed in the presence of a transcription inhibitor while it is completely blocked in the presence of a translation inhibitor (for review, see [59]). This means that stored mRNAs, also called ‘long-lived mRNAs’, are necessary and sufficient to carry out the germination in sensu stricto which corresponds to the determining phase of dormancy maintenance or alleviation.
Non-enzymatic oxidations are possible in low hydrated seeds and represent the most plausible lead to explain the observed molecular changes reported during after-ripening [60]. Indeed, mRNA oxidation was shown to be associated with dormancy release during after-ripening in sunflower and wheat [61,62], which alters the stability of stored mRNAs, being finally degraded or translated into non-functional proteins [63]. However, if a fraction of stored mRNA is inactivated, the one involved in germination has to be protected from oxidation. A recent study showed that the association of mRNA with monosomes may be the key process for mRNA preservation [64]. The identification of translated proteins from stored mRNA in rice seeds showed that they correspond to glycolysis and translation machinery, and newly synthetized mRNA are involved in pyruvate metabolism, tricarboxylic acid (TCA) cycle, or momilactone biosynthesis [65]. This indicates that these newly synthetized energy components may represent good candidates for the regulation of germination. In fact, it was shown that TCA enzyme regulation participates in the control of seed dormancy in sunflower [16]. It was also shown that TCA enzymes were thiol redox regulated and responsible for efficient TCA functioning [41]. Several other post-translational modifications, such as phosphorylation, ubiquitination, carbonylation, glycosylation, acetylation, succinylation, or sumoylation, have been proposed to play important roles in seed germination by controlling hormonal signaling, metabolism, and redox status (for review, see [66]). Carbonylation represents the most plausible modification that takes place at dry state as a consequence of the non-enzymatic generation of reactive oxygen and nitrogen species [67]. It has been shown that protein carbonylation occurs during after-ripening and may play an important role in the transition from dormant to non-dormant state in dry seeds by facilitating reserve degradation and regulating cell signaling [68].
Respiration and redox regulation therefore constitute the most important regulation in the initiation of the germination process, but hormonal regulation that takes place later during imbibition is also crucial to germination achievement.

3.2. Internal Determinants of Dormancy

It is well established that dormancy is regulated by the hormonal balance between the main positive regulator abscisic acid (ABA) and negative ones, such as hormones like gibberellic acid (GA), ethylene (ET), auxins, or brassinosteroids, as well as some other molecules, like ROS or nitric oxide (NO). The involvement of each of them and their interactions in the whole process of germination depend on the structure of the seed and the environment. Nevertheless, ABA represents the highly conserved component of the process across species and the unique dormancy determinant as opposed to the multiple stimulants of germination. Thus, to illustrate the regulation and function of hormones in the physiology of germination, without elaborating on all the hormones and their complex signaling, the case of the ABA is the most appropriate. High ABA is induced during the maturation phase of seed development to set up desiccation and dormancy. In mature seeds, a large proportion of the stored mRNA was shown to correspond to genes in which the promoters are targets of ABA-responsive transcription factors, which could be the residual consequence of the ABA induction in the maturation phase or a regulated process to insure the execution of ABA signaling upon imbibition [69]. Upon imbibition, ABA content declines similarly in ND and D seeds during the early phase of germination, but this decrease continues in ND seeds while subsequent de novo ABA synthesis occurred in imbibed D seeds, leading to dormancy maintenance [70]. Therefore, ABA biosynthesis, catalysis enzymes, and corresponding genes, nine-cis-epoxycarotenoid dioxygenase (NCED) and cytochrome P450 707A (CYP707A), respectively, represent the major determinants of seed dormancy. However, a decrease in ABA content is not a prerequisite for germination as ABA signaling events represent another level of regulation [71]. The responsiveness of seeds to ABA is called ABA sensitivity and it involves several promoters, genes, and protein regulations. In recent studies, a number of these key players have been characterized in a complex network partly connected with other hormones having a dual role in this process [72,73]. However, how such players operate to arrest expansion of the embryo and growth remains unsolved. Considering the challenging energy demand of the germination process, reserve breakdown and respiration may represent the regulatory mechanisms. ABA treatment is able to inhibit reserve mobilization and sugar treatment to overcome the exogenous ABA inhibition of germination. However, the effect of endogenous ABA is still unclear [74]. Mitochondria play a central role in energy supply and they are also associated with ABA sensitivity based on works showing that several mutants of RNA processing for subunits of the electron chain display reduced ABA sensitivity. This regulation involves retrograde, anterograde, and inter-organelle signals in the transcription control of the ABA biosynthesis gene, NCED [75]. On the other hand, Paszkiewicz et al. [10] have shown that mitochondrial dynamics associated with germination condition was slightly affected by ABA treatment, arguing that mitochondrion reactivation depends only on the physical conditions of hydration and temperature. Based on these works, the optimal differentiation and functioning of mitochondrion are associated with an ABA sensitivity decrease. Accordingly, it is easy to consider that in dormant seeds, the impairment of mitochondrial activity occurs. However, it has long been established that inhibitors of oxidative phosphorylation such as cyanid can break dormancy. This paradox has still not been elucidated. The activation of the pentose phosphate pathway, the metabolic pathway that supplies reducing energy to cells, has been the most plausible hypothesis proposed [76]. Indeed, in reduced mitochondrial activity, glycolysis is activated to obtain ATP, a phenomenon known the “Pasteur effect”, leading to pyruvate production and the accumulation of fermentation by-products. Thus, anaerobic metabolism facilitates reserve breakdown and it might operate, in normal conditions, at the onset of germination when the mitochondria are not yet fully reactivated. All these data point to the importance of cell metabolism and energy regulation for successful germination.
On the other hand, it was proposed that the ABA inhibition of growth in germinating Arabidopsis seeds is driven by its inhibitory action on PM H+-ATPase activity [77]. ABA inhibition was less effective in Arabidopsis mutants with increased capacity for H+ efflux, suggesting that cytosolic acidification due to reduced H+-ATPase activity was the main mechanism driving growth inhibition [77]. Similarly, in sunflower, ABA induced the inhibition of PM H+-ATPase in non-dormant seeds, which display hyperpolarization and subsequent membrane energization. Meanwhile, in dormant seeds, PM H+-ATPase activity was reduced even if the corresponding proteins were present and the levels of ATP were comparable to that on ND [52]. PM H+-ATPase activity is also regulated by ROS and ethylene in the opposite way [52], as proposed in the model presented Figure 4. Moreover, ND cell hyperpolarization allows sugar influx through H+/sugar symporter [52] and it was shown that glucose and fructose contents were higher in ND as compared to D seeds in the same seed model [17]. Such sugar influx along with other solutes such as K+ driven by the proton motive force of PM H+-ATPase activity could be the prerequisite for dormancy alleviation and germination as membrane energization represents the starting point for metabolism resumption by influencing water and hexose movement, but also ions and particularly protons influencing mitochondrial activity.
At the scale of the whole organ, specialized tissues play a critical role upon imbibition. The most striking events are the transport of hormones from/to the different tissues of the seed and consequently their specific contribution in each tissue to regulating seed germination. For example, in endospermic seeds, ABA is produced in the endosperm and transported to the embryo, while GA goes in the opposite way [78]. Moreover, cell wall loosening and programmed cell death occur specifically in the endosperm to facilitate root protrusion [79]. In non-endospermic seeds, very scarce information is available but a recent study has shown a differentiated localization of ABA, GA, and ethylene in the meristematic zone as compared to the other parts of the seed [71]. How such tissues are differentially programmed to fulfill their respective roles and what biological structure or genetic program confers them their ability, are yet to be discovered. However, whatever the state of the seed, the signal by which it awakens comes from the environment.

3.3. Environmental Impact on Dormancy

Environmental factors are of high importance in the awakening of the seed. Their effect on seed performance was shown to surpass genetic impact [80]. Temperature and soil moisture oscillations are the major players under natural conditions. Indeed, alternating temperatures more than constant ones can promote germination via the interplay between ROS signaling and hormones [81]. In fact, it has been shown that fluctuating temperatures alleviate dormancy by reducing ABA synthesis and signaling [82]. In the absence of a change in ABA content, there was a decrease in ABA sensitivity in sunflower dormant seeds in response to constant temperature that induced dormancy alleviation [71]. The activity of several enzymes of TCA and glycolysis were shown to be altered in the same model [16]. In wheat, high temperature treatment during seed development affects mitochondrion functioning by reducing the respiration rate and ATP content [83]. Indeed, elevated temperatures experienced by the mother plant during seed development and maturation had a negative effect on seed composition, germination, and vigor (for review, see [84]). On the other hand, light and nitrate also play important roles. Their effects are associated in dormancy cycling [85,86,87]. Seed sensitivity to both of them depends on the season and depth of dormancy. A low concentration of nitrate (around 0.1 mM) is able to promote seed germination in several species [85]. Several evidences converge towards nitrate induction of CYP707A2 leading to ABA decrease more than GA biosynthesis in dormancy breakdown [85]. However, GA biosynthesis gene involvement has also been reported in response to environmental cues in Arabidopsis seeds from lab but also soil seed bank experiences [88,89]. Moreover, the analysis of the whole transcriptome change by nitrate treatment during seed imbibition showed the upregulation of genes involved in nitrate assimilation and transport, hormone metabolism, and energy, such as Glucose-6-phosphate dehydrogenase2, highlighting the importance of the pentose phosphate pathway [90]. Interestingly, at the level of gene expression, different environmental signals, such as light, nitrate, stratification, or after-ripening induced common changes associated to dormancy release [88]. They concern genes belonging to translation machinery, cell wall modification, and reserve mobilization. Such changes in transcript abundance are reversible, allowing the dormancy cycling phenomenon which occurs on entering a secondary dormancy when unfavorable environmental conditions are prolonged after the primary dormancy alleviation. In fact, dormancy is tightly regulated in natural conditions as in the soil, when seeds experience several scenarios of temperature, light, nitrate, and moisture, as well as microbial environment. The latter corresponds to a wide range of microbes, as pathogenic ones can induce decrease in seed longevity due to infection, and others can influence seed dormancy by breaking down the seed coat [12,91].
Thus, environmental cues influence not only seed dormancy alleviation, but determine the depth of dormancy mediated by the mother plant during seed development and maturation. The understanding of such influence is crucial for agriculture, especially in the context of environmental condition fluctuations due to global warming.

4. Seeds: The Ability to Recover from Ageing

Seeds remain viable, i.e., capable to germinate producing a viable plantlet, generally for a long period, from weeks to thousands of years, depending on species. Seed longevity is important for economic aspects of trade and agronomy associated with storage, but also for maintaining biodiversity. Obviously, reduced water content and metabolic activity enable such great longevity. However, long-term conservation results in a loss of viability due to deterioration processes. More specifically, unsuitable conditions of conservation, such as temperature and moisture, experienced by seeds in natural conditions or during storage accelerate deterioration processes, such as the loss of membrane integrity and oxidation of macromolecules, leading to the impairment of metabolism. However, the seed has the extraordinary ability to recover using extensive repair machinery, which represents another performance to stay alive.

4.1. Seed Ageing

Seed life span is of importance for field crop species impacting agriculture, but also for plant species diversity maintenance by its impact on seed longevity in the soil [92]. Seed ageing was defined as the loss of seed quality and viability over time [93]. Aged seeds germinate poorly giving abnormal seedlings or ultimately are unable to germinate. Orthodox seeds are resistant to ageing for very long time because they have a very low water content, resulting in reduced cell metabolism especially respiration which is responsible for the major production of ROS [94]. Indeed, as for the dry after-ripening process described above, enzymatic reactions and respiration are restricted by the lack of free available water preventing cellular damage. It was shown that long term storage resulting in seed loss of viability is associated with the impairment of mitochondrial activity and protein synthesis machinery [95,96]. Mitochondrion membrane integrity was identified as the primary target for ageing leading to the deregulation of its oxidative properties [97]. ROS are considered as the major cause of seed deterioration due to the oxidation of its components [98]. Lipid peroxidation has been reported in several studies on different seed species, influencing lipid metabolism and membrane integrity, as for sunflower seeds [99,100]. Furthermore, DNA damage producing double- or single- strand breaks or damaged bases is responsible for genome integrity loss and subsequent low seed quality [101]. DNA laddering has been shown in sunflower and pea aged seeds when damage extent exceeded repair capacity, pointing out the key role of mitochondria dysfunction in seed ageing [102,103]. Indeed, a direct correlation between ROS production and mitochondrial impairment leading to programed cell death has been shown in Ulmus pumila L. [104]. At last, total RNA content and integrity was shown to decrease in aged soybean seeds, with greater resistance of the shortest transcript (<1200 bp) mainly involved in ribosomal and translational functions, as compared to the longer transcripts (>2500 bp) corresponding to proteins with ATP-binding functions, indicating that stored mRNAs may be involved in seed longevity [105,106].
If cellular damage is inevitably induced during ageing, the seed can resist using several features. In fact, at the cellular level, several components protect from cell damage, such as sugars, LEA, dehydrins, or heat shock proteins involved in dehydration–rehydration protection or storage proteins being preferentially oxidized protecting vital cell components from ROS damage [92]. Indeed, in the soil, seeds experience changes in temperature and water content, two major factors that influence biochemical reaction resumption, inducing ROS production and associated damages. In addition to cellular organization, the efficacy of dormancy in preventing growth resumption and the ability for damage repair are mechanisms of importance for seed longevity in natural conditions. Such repair processes are involved in the extraordinary ability of the seed to recover from ageing. They have been explored during seed priming treatment and are of interest for all living organisms.

4.2. Seed Priming

Seeds possess effective repair machinery to cope with ageing-associated oxidative damage. Several non-enzymatic antioxidants have been proposed to be determinant in seed longevity, e.g., glutathione or ascorbic acid [107,108]. The antioxidant enzymes are also of importance in ROS detoxification, such as catalase, superoxide dismutase, ascorbate peroxidase, or glutathione reductase [109,110]. Other enzymes acting on specific macromolecules are also activated, such as DNA or protein repair enzymes [101,111,112]. Such machinery operates when seed hydration occurred, and its efficiency depends on the plant species and the extent of ageing damage. Based on this feature, a priming technique has been developed to improve seed quality from alterations caused by several stresses. Priming treatment consists of seed pre-hydration with a controlled amount of water which does not allow radicle elongation, i.e., a water amount corresponding phase II of the germination sensu stricto, which is sufficient to trigger the reparation processes. The addition of beneficial molecules, such as antioxidants or hormones, during priming treatment can further increase seed reparation and subsequent quality. Several priming techniques have been developed depending on the plant species and subsequent use. They all lead to an improvement of seed performance under variable environmental conditions [113]. It was shown that repair mechanisms and oxidative management in primed seeds represent the main processes associated with priming induced germination improvement, but DNA replication, cell cycle advancement, the modification of the membrane structure, and restoration of mitochondrial integrity were also proposed to explain the priming effect in germination improvement [114]. In fact, the seed engages growth preparation processes during imbibition while maintaining the desiccation tolerance machinery which allowed a successful dehydration after the priming treatment. The dried primed seed is ready to grow better, even under stress conditions. This feature is interesting for agriculture, which is why the priming technique is widely used. Priming may be even more useful in the years to come due to changing climate conditions.

5. Conclusions

Understanding the successful entry and exit from desiccation is fundamental for the improvement of seed germination under challenging conditions anticipated due to global warming. The application in plant germplasm conservation in seed banks is of high importance in the maintenance of genetic resources for food and environment security. In this review, several layers of regulations of seed performance were shown, from the organization and physical protection of cell components to the regulation of several signaling processes in a coordinated crosstalk. However, their implementation and the coordination of these mechanisms during seed development deserve more investigations.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not associated data marked.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Meimoun, P.; Mordret, E.; Langlade, N.B.; Balzergue, S.; Arribat, S.; Bailly, C.; El-Maarouf-Bouteau, H. Is Gene Transcription Involved in Seed Dry After-Ripening? PLoS ONE 2014, 9, e86442. [Google Scholar] [CrossRef]
  2. Roberts, E.H. Predicting the Storage Life of Seeds. Seed Sci. Technol. 1973, 1, 499–514. [Google Scholar]
  3. Ellis, R.H.; Hong, T.D.; Roberts, E.H. An Intermediate Category of Seed Storage Behaviour? I. Coffee. J. Exp. Bot. 1990, 41, 1167–1174. [Google Scholar] [CrossRef]
  4. Benner, S.A. Defining Life. Astrobiology 2010, 10, 1021–1030. [Google Scholar] [CrossRef] [Green Version]
  5. Vitas, M.; Dobovišek, A. In the Beginning Was a Mutualism—On the Origin of Translation. Orig. Life Evol. Biospheres 2018, 48, 223–243. [Google Scholar] [CrossRef]
  6. Belmonte, M.F.; Kirkbride, R.C.; Stone, S.L.; Pelletier, J.M.; Bui, A.Q.; Yeung, E.C.; Hashimoto, M.; Fei, J.; Harada, C.M.; Munoz, M.D.; et al. Comprehensive Developmental Profiles of Gene Activity in Regions and Subregions of the Arabidopsis Seed. Proc. Natl. Acad. Sci. USA 2013, 110, E435–E444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Ingram, G.C. Family Life at Close Quarters: Communication and Constraint in Angiosperm Seed Development. Protoplasma 2010, 247, 195–214. [Google Scholar] [CrossRef]
  8. Ballesteros, D.; Walters, C. Detailed Characterization of Mechanical Properties and Molecular Mobility within Dry Seed Glasses: Relevance to the Physiology of Dry Biological Systems: Molecular Mobility within the Glass of Dry Seed. Plant J. 2011, 68, 607–619. [Google Scholar] [CrossRef] [PubMed]
  9. Vertucci, C.W.; Leopold, A.C. Bound Water in Soybean Seed and Its Relation to Respiration and Imbibitional Damage. Plant Physiol. 1984, 75, 114–117. [Google Scholar] [CrossRef] [PubMed]
  10. Paszkiewicz, G.; Gualberto, J.M.; Benamar, A.; Macherel, D.; Logan, D.C. Arabidopsis Seed Mitochondria Are Bioenergetically Active Immediately upon Imbibition and Specialize via Biogenesis in Preparation for Autotrophic Growth. Plant Cell 2017, 29, 109–128. [Google Scholar] [CrossRef]
  11. Sallon, S.; Solowey, E.; Cohen, Y.; Korchinsky, R.; Egli, M.; Woodhatch, I.; Simchoni, O.; Kislev, M. Germination, Genetics, and Growth of an Ancient Date Seed. Science 2008, 320, 1464. [Google Scholar] [CrossRef]
  12. Leprince, O.; Pellizzaro, A.; Berriri, S.; Buitink, J. Late Seed Maturation: Drying without Dying. J. Exp. Bot. 2017, 68, 827–841. [Google Scholar] [CrossRef] [Green Version]
  13. Angelovici, R.; Galili, G.; Fernie, A.R.; Fait, A. Seed Desiccation: A Bridge between Maturation and Germination. Trends Plant Sci. 2010, 15, 211–218. [Google Scholar] [CrossRef] [PubMed]
  14. Han, C.; Zhen, S.; Zhu, G.; Bian, Y.; Yan, Y. Comparative Metabolome Analysis of Wheat Embryo and Endosperm Reveals the Dynamic Changes of Metabolites during Seed Germination. Plant Physiol. Biochem. 2017, 115, 320–327. [Google Scholar] [CrossRef] [PubMed]
  15. Howell, K.A.; Narsai, R.; Carroll, A.; Ivanova, A.; Lohse, M.; Usadel, B.; Millar, A.H.; Whelan, J. Mapping Metabolic and Transcript Temporal Switches during Germination in Rice Highlights Specific Transcription Factors and the Role of RNA Instability in the Germination Process. Plant Physiol. 2009, 149, 961–980. [Google Scholar] [CrossRef] [Green Version]
  16. Xia, Q.; Ponnaiah, M.; Cueff, G.; Rajjou, L.; Prodhomme, D.; Gibon, Y.; Bailly, C.; Corbineau, F.; Meimoun, P.; El-Maarouf-Bouteau, H. Integrating Proteomics and Enzymatic Profiling to Decipher Seed Metabolism Affected by Temperature in Seed Dormancy and Germination. Plant Sci. Int. J. Exp. Plant Biol. 2018, 269, 118–125. [Google Scholar] [CrossRef]
  17. Xia, Q.; Saux, M.; Ponnaiah, M.; Gilard, F.; Perreau, F.; Huguet, S.; Balzergue, S.; Langlade, N.; Bailly, C.; Meimoun, P.; et al. One Way to Achieve Germination: Common Molecular Mechanism Induced by Ethylene and After-Ripening in Sunflower Seeds. Int. J. Mol. Sci. 2018, 19, 2464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Weitbrecht, K.; Müller, K.; Leubner-Metzger, G. First off the Mark: Early Seed Germination. J. Exp. Bot. 2011, 62, 3289–3309. [Google Scholar] [CrossRef] [Green Version]
  19. Nonogaki, H.; Bassel, G.W.; Bewley, J.D. Germination—Still a Mystery. Plant Sci. 2010, 179, 574–581. [Google Scholar] [CrossRef]
  20. Szent-Györgyi, A. Cell-Associated Water; Academic Press: New York, NY, USA, 1979; pp. 363–413. [Google Scholar]
  21. Ball, P. Water as an Active Constituent in Cell Biology. Chem. Rev. 2008, 108, 74–108. [Google Scholar] [CrossRef]
  22. Bellissent-Funel, M.-C.; Hassanali, A.; Havenith, M.; Henchman, R.; Pohl, P.; Sterpone, F.; van der Spoel, D.; Xu, Y.; Garcia, A.E. Water Determines the Structure and Dynamics of Proteins. Chem. Rev. 2016, 116, 7673–7697. [Google Scholar] [CrossRef] [PubMed]
  23. Vertucci, C.W. Calorimetric Studies of the State of Water in Seed Tissues. Biophys. J. 1990, 58, 1463–1471. [Google Scholar] [CrossRef] [Green Version]
  24. Buitink, J.; Leprince, O. Intracellular Glasses and Seed Survival in the Dry State. Comptes R. Biol. 2008, 331, 788–795. [Google Scholar] [CrossRef] [PubMed]
  25. Rupley, J.A.; Gratton, E.; Careri, G. Water and Globular Proteins. Trends Biochem. Sci. 1983, 8, 18–22. [Google Scholar] [CrossRef] [Green Version]
  26. Richards, F.M.; Richmond, T. Solvents, Interfaces and Protein Structure. In Novartis Foundation Symposia; Porter, R., Fitzsimons, D.W., Eds.; John Wiley & Sons, Ltd.: Chichester, UK, 2008; pp. 23–45. ISBN 978-0-470-72042-4. [Google Scholar]
  27. Texter, J. Nucleic Acid-Water Interactions. Prog. Biophys. Mol. Biol. 1979, 33, 83–97. [Google Scholar] [CrossRef]
  28. Shweta, H.; Sen, S. Dynamics of Water and Ions around DNA: What Is so Special about Them? J. Biosci. 2018, 43, 499–518. [Google Scholar] [CrossRef]
  29. van Zanten, M.; Koini, M.A.; Geyer, R.; Liu, Y.; Brambilla, V.; Bartels, D.; Koornneef, M.; Fransz, P.; Soppe, W.J.J. Seed Maturation in Arabidopsis Thaliana Is Characterized by Nuclear Size Reduction and Increased Chromatin Condensation. Proc. Natl. Acad. Sci. USA 2011, 108, 20219–20224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Borisjuk, L.; Rolletschek, H. The Oxygen Status of the Developing Seed: Tansley Review. New Phytol. 2009, 182, 17–30. [Google Scholar] [CrossRef]
  31. Logan, D.C.; Millar, A.H.; Sweetlove, L.J.; Hill, S.A.; Leaver, C.J. Mitochondrial Biogenesis during Germination in Maize Embryos. Plant Physiol. 2001, 125, 662–672. [Google Scholar] [CrossRef] [Green Version]
  32. Howell, K.A.; Millar, A.H.; Whelan, J. Ordered Assembly of Mitochondria During Rice Germination Begins with Promitochondrial Structures Rich in Components of the Protein Import Apparatus. Plant Mol. Biol. 2006, 60, 201–223. [Google Scholar] [CrossRef]
  33. Nawa, Y.; Asahi, T. Rapid Development of Mitochondria in Pea Cotyledons during the Early Stage of Germination. Plant Physiol. 1971, 48, 671–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Morohashi, Y.; Bewley, J.D. Development of Mitochondrial Activities in Pea Cotyledons: Influence of desiccation during and following germination of the axis. Plant Physiol. 1980, 66, 637–640. [Google Scholar] [CrossRef] [Green Version]
  35. Morohashi, Y.; Bewley, J.D.; Yeung, E.C. Biogenesis of Mitochondria in Imbibed Peanut Cotyledons: II. Development of light and heavy mitochondria. Plant Physiol. 1981, 68, 318–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Ehrenshaft, M.; Brambl, R. Respiration and Mitochondrial Biogenesis in Germinating Embryos of Maize. Plant Physiol. 1990, 93, 295–304. [Google Scholar] [CrossRef] [Green Version]
  37. Law, S.R.; Narsai, R.; Whelan, J. Mitochondrial Biogenesis in Plants during Seed Germination. Mitochondrion 2014, 19, 214–221. [Google Scholar] [CrossRef]
  38. Czarna, M.; Kolodziejczak, M.; Janska, H. Mitochondrial Proteome Studies in Seeds during Germination. Proteomes 2016, 4, 19. [Google Scholar] [CrossRef]
  39. Benamar, A.; Rolletschek, H.; Borisjuk, L.; Avelange-Macherel, M.-H.; Curien, G.; Mostefai, H.A.; Andriantsitohaina, R.; Macherel, D. Nitrite–Nitric Oxide Control of Mitochondrial Respiration at the Frontier of Anoxia. Biochim. Biophys. Acta BBA—Bioenerg. 2008, 1777, 1268–1275. [Google Scholar] [CrossRef]
  40. Attucci, S.; Carde, J.P.; Raymond, P.; Saint-Gès, V.; Spiteri, A.; Pradet, A. Oxidative Phosphorylation by Mitochondria Extracted from Dry Sunflower Seeds. Plant Physiol. 1991, 95, 390–398. [Google Scholar] [CrossRef] [PubMed]
  41. Nietzel, T.; Mostertz, J.; Ruberti, C.; Née, G.; Fuchs, P.; Wagner, S.; Moseler, A.; Müller-Schüssele, S.J.; Benamar, A.; Poschet, G.; et al. Redox-Mediated Kick-Start of Mitochondrial Energy Metabolism Drives Resource-Efficient Seed Germination. Proc. Natl. Acad. Sci. USA 2020, 117, 741–751. [Google Scholar] [CrossRef]
  42. Buchanan, B.B. The Path to Thioredoxin and Redox Regulation Beyond Chloroplasts. Plant Cell Physiol. 2017, 58, 1826–1832. [Google Scholar] [CrossRef] [PubMed]
  43. Simon, E.W. Phospholipids and plant membrane permeability. New Phytol. 1974, 73, 377–420. [Google Scholar] [CrossRef]
  44. Yu, X.; Li, A.; Li, W. How Membranes Organize during Seed Germination: Three Patterns of Dynamic Lipid Remodelling Define Chilling Resistance and Affect Plastid Biogenesis: Remodelling of Membrane Lipids during Germination. Plant Cell Environ. 2015, 38, 1391–1403. [Google Scholar] [CrossRef] [PubMed]
  45. Lin, Y.; Xin, X.; Yin, G.; He, J.; Zhou, Y.; Chen, J.; Lu, X. Membrane Phospholipids Remodeling upon Imbibition in Brassica Napus L. Seeds. Biochem. Biophys. Res. Commun. 2019, 515, 289–295. [Google Scholar] [CrossRef] [PubMed]
  46. Sze, H.; Li, X.; Palmgren, M.G. Energization of Plant Cell Membranes by H+-Pumping ATPases: Regulation and Biosynthesis. Plant Cell 1999, 11, 677–689. [Google Scholar] [CrossRef] [Green Version]
  47. Haruta, M.; Burch, H.L.; Nelson, R.B.; Barrett-Wilt, G.; Kline, K.G.; Mohsin, S.B.; Young, J.C.; Otegui, M.S.; Sussman, M.R. Molecular Characterization of Mutant Arabidopsis Plants with Reduced Plasma Membrane Proton Pump Activity. J. Biol. Chem. 2010, 285, 17918–17929. [Google Scholar] [CrossRef] [Green Version]
  48. Falhof, J.; Pedersen, J.T.; Fuglsang, A.T.; Palmgren, M. Plasma Membrane H+-ATPase Regulation in the Center of Plant Physiology. Mol. Plant 2016, 9, 323–337. [Google Scholar] [CrossRef] [Green Version]
  49. Lang, V.; Pertl-Obermeyer, H.; Safiarian, M.J.; Obermeyer, G. Pump up the Volume—A Central Role for the Plasma Membrane H+ Pump in Pollen Germination and Tube Growth. Protoplasma 2014, 251, 477–488. [Google Scholar] [CrossRef]
  50. Pedersen, J.T.; Falhof, J.; Ekberg, K.; Buch-Pedersen, M.J.; Palmgren, M. Metal Fluoride Inhibition of a P-Type H+ Pump: Stabilization of the phosphoenzyme intermediate contributes to post-translational pump activation. J. Biol. Chem. 2015, 290, 20396–20406. [Google Scholar] [CrossRef] [Green Version]
  51. Pedersen, J.T.; Kanashova, T.; Dittmar, G.; Palmgren, M. Isolation of Native Plasma Membrane H+-ATP Ase (Pma1p) in Both the Active and Basal Activation States. FEBS Open Bio 2018, 8, 774–783. [Google Scholar] [CrossRef] [Green Version]
  52. De Bont, L.; Naim, E.; Arbelet-Bonnin, D.; Xia, Q.; Palm, E.; Meimoun, P.; Mancuso, S.; El-Maarouf-Bouteau, H.; Bouteau, F. Activation of Plasma Membrane H+-ATPases Participates in Dormancy Alleviation in Sunflower Seeds. Plant Sci. 2019, 280, 408–415. [Google Scholar] [CrossRef] [Green Version]
  53. Leubner-Metzger, G. Beta-1,3-Glucanase Gene Expression in Low-Hydrated Seeds as a Mechanism for Dormancy Release during Tobacco after-Ripening. Plant J. Cell Mol. Biol. 2005, 41, 133–145. [Google Scholar] [CrossRef]
  54. Bove, J.; Lucas, P.; Godin, B.; Ogé, L.; Jullien, M.; Grappin, P. Gene Expression Analysis by CDNA-AFLP Highlights a Set of New Signaling Networks and Translational Control during Seed Dormancy Breaking in Nicotiana Plumbaginifolia. Plant Mol. Biol. 2005, 57, 593–612. [Google Scholar] [CrossRef] [PubMed]
  55. Leymarie, J.; Bruneaux, E.; Gibot-Leclerc, S.; Corbineau, F. Identification of Transcripts Potentially Involved in Barley Seed Germination and Dormancy Using CDNA-AFLP. J. Exp. Bot. 2006, 58, 425–437. [Google Scholar] [CrossRef] [Green Version]
  56. Gao, F.; Jordan, M.C.; Ayele, B.T. Transcriptional Programs Regulating Seed Dormancy and Its Release by After-Ripening in Common Wheat (Triticum Aestivum L.). Plant Biotechnol. J. 2012, 10, 465–476. [Google Scholar] [CrossRef] [PubMed]
  57. Liu, A.; Gao, F.; Kanno, Y.; Jordan, M.C.; Kamiya, Y.; Seo, M.; Ayele, B.T. Regulation of Wheat Seed Dormancy by After-Ripening Is Mediated by Specific Transcriptional Switches That Induce Changes in Seed Hormone Metabolism and Signaling. PLoS ONE 2013, 8, e56570. [Google Scholar] [CrossRef] [Green Version]
  58. Dure, L.; Waters, L. Long-lived messenger RNA: Evidence from cotton seed germination. Science 1965, 147, 410–412. [Google Scholar] [CrossRef]
  59. Sano, N.; Rajjou, L.; North, H.M. Lost in Translation: Physiological Roles of Stored MRNAs in Seed Germination. Plants 2020, 9, 347. [Google Scholar] [CrossRef] [Green Version]
  60. El-Maarouf-Bouteau, H.; Bailly, C. Oxidative Signaling in Seed Germination and Dormancy. Plant Signal. Behav. 2008, 3, 175–182. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Bazin, J.; Langlade, N.; Vincourt, P.; Arribat, S.; Balzergue, S.; El-Maarouf-Bouteau, H.; Bailly, C. Targeted MRNA Oxidation Regulates Sunflower Seed Dormancy Alleviation during Dry After-Ripening. Plant Cell 2011, 23, 2196–2208. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Gao, F.; Rampitsch, C.; Chitnis, V.R.; Humphreys, G.D.; Jordan, M.C.; Ayele, B.T. Integrated Analysis of Seed Proteome and MRNA Oxidation Reveals Distinct Post-Transcriptional Features Regulating Dormancy in Wheat (Triticum Aestivum L.). Plant Biotechnol. J. 2013, 11, 921–932. [Google Scholar] [CrossRef]
  63. Katsuya-Gaviria, K.; Caro, E.; Carrillo-Barral, N.; Iglesias-Fernández, R. Reactive Oxygen Species (ROS) and Nucleic Acid Modifications during Seed Dormancy. Plants 2020, 9, 679. [Google Scholar] [CrossRef] [PubMed]
  64. Buijs, G.; Vogelzang, A.; Nijveen, H.; Bentsink, L. Dormancy Cycling: Translation-related Transcripts Are the Main Difference between Dormant and Non-dormant Seeds in the Field. Plant J. 2020, 102, 327–339. [Google Scholar] [CrossRef] [PubMed]
  65. Sano, N.; Takebayashi, Y.; To, A.; Mhiri, C.; Rajjou, L.; Nakagami, H.; Kanekatsu, M. Shotgun Proteomic Analysis Highlights the Roles of Long-Lived MRNAs and De Novo Transcribed MRNAs in Rice Seeds upon Imbibition. Plant Cell Physiol. 2019, 60, 2584–2596. [Google Scholar] [CrossRef] [PubMed]
  66. Yu, F.; Li, M.; He, D.; Yang, P. Advances on Post-Translational Modifications Involved in Seed Germination. Front. Plant Sci. 2021, 12, 642979. [Google Scholar] [CrossRef]
  67. Arc, E.; Galland, M.; Cueff, G.; Godin, B.; Lounifi, I.; Job, D.; Rajjou, L. Reboot the System Thanks to Protein Post-Translational Modifications and Proteome Diversity: How Quiescent Seeds Restart Their Metabolism to Prepare Seedling Establishment. PROTEOMICS 2011, 11, 1606–1618. [Google Scholar] [CrossRef]
  68. Oracz, K.; El-Maarouf Bouteau, H.; Farrant, J.M.; Cooper, K.; Belghazi, M.; Job, C.; Job, D.; Corbineau, F.; Bailly, C. ROS Production and Protein Oxidation as a Novel Mechanism for Seed Dormancy Alleviation. Plant J. Cell Mol. Biol. 2007, 50, 452–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Nakabayashi, K.; Okamoto, M.; Koshiba, T.; Kamiya, Y.; Nambara, E. Genome-Wide Profiling of Stored MRNA in Arabidopsis Thaliana Seed Germination: Epigenetic and Genetic Regulation of Transcription in Seed. Plant J. Cell Mol. Biol. 2005, 41, 697–709. [Google Scholar] [CrossRef]
  70. Nambara, E.; Okamoto, M.; Tatematsu, K.; Yano, R.; Seo, M.; Kamiya, Y. Abscisic Acid and the Control of Seed Dormancy and Germination. Seed Sci. Res. 2010, 20, 55–67. [Google Scholar] [CrossRef]
  71. Xia, Q.; Ponnaiah, M.; Thanikathansubramanian, K.; Corbineau, F.; Bailly, C.; Nambara, E.; Meimoun, P.; El-Maarouf-Bouteau, H. Re-Localization of Hormone Effectors Is Associated with Dormancy Alleviation by Temperature and after-Ripening in Sunflower Seeds. Sci. Rep. 2019, 9, 4861. [Google Scholar] [CrossRef] [Green Version]
  72. Shu, K.; Liu, X.; Xie, Q.; He, Z. Two Faces of One Seed: Hormonal Regulation of Dormancy and Germination. Mol. Plant 2016, 9, 34–45. [Google Scholar] [CrossRef] [Green Version]
  73. Ali, F.; Qanmber, G.; Li, F.; Wang, Z. Updated Role of ABA in Seed Maturation, Dormancy, and Germination. J. Adv. Res. 2022, 35, 199–214. [Google Scholar] [CrossRef] [PubMed]
  74. Finkelstein, R.R.; Gampala, S.S.L.; Rock, C.D. Abscisic Acid Signaling in Seeds and Seedlings. Plant Cell 2002, 14 (Suppl. 1), S15–S45. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Nonogaki, H. The Long-Standing Paradox of Seed Dormancy Unfolded? Trends Plant Sci. 2019, 24, 989–998. [Google Scholar] [CrossRef] [PubMed]
  76. Bewley, J.D.; Black, M. Seeds; Springer: Boston, MA, USA, 1994; pp. 1–33. Available online: https://0-link-springer-com.brum.beds.ac.uk/chapter/10.1007/978-1-4899-1002-8_1 (accessed on 15 October 2021).
  77. Planes, M.D.; Niñoles, R.; Rubio, L.; Bissoli, G.; Bueso, E.; García-Sánchez, M.J.; Alejandro, S.; Gonzalez-Guzmán, M.; Hedrich, R.; Rodriguez, P.L.; et al. A Mechanism of Growth Inhibition by Abscisic Acid in Germinating Seeds of Arabidopsis thaliana Based on Inhibition of Plasma Membrane H+-ATPase and Decreased Cytosolic PH, K+, and Anions. J. Exp. Bot. 2015, 66, 813–825. [Google Scholar] [CrossRef] [Green Version]
  78. Penfield, S. Seed Dormancy and Germination. Curr. Biol. 2017, 27, R874–R878. [Google Scholar] [CrossRef] [Green Version]
  79. Yan, D.; Duermeyer, L.; Leoveanu, C.; Nambara, E. The Functions of the Endosperm During Seed Germination. Plant Cell Physiol. 2014, 55, 1521–1533. [Google Scholar] [CrossRef] [PubMed]
  80. de Souza Vidigal, D.; He, H.; Hilhorst, H.W.M.; Willems, L.A.J.; Bentsink, L. Arabidopsis in the Wild—The Effect of Seasons on Seed Performance. Plants 2020, 9, 576. [Google Scholar] [CrossRef]
  81. Matilla, A.J. Seed Dormancy: Molecular Control of Its Induction and Alleviation. Plants 2020, 9, 1402. [Google Scholar] [CrossRef]
  82. Roberto Huarte, H.; Luna, V.; Pagano, E.A.; Zavala, J.A.; Benech-Arnold, R.L. Fluctuating Temperatures Terminate Dormancy in Cynara cardunculus Seeds by Turning off ABA Synthesis and Reducing ABA Signalling, but Not Stimulating GA Synthesis or Signalling. Seed Sci. Res. 2014, 24, 79–89. [Google Scholar] [CrossRef] [Green Version]
  83. Grass, L.; Burris, J.S. Effect of Heat Stress during Seed Development and Maturation on Wheat (Triticum durum) Seed Quality. I. Seed Germination and Seedling Vigor. Can. J. Plant Sci. 1995, 75, 821–829. [Google Scholar] [CrossRef] [Green Version]
  84. Domergue, J.; Abadie, C.; Limami, A.; Way, D.; Tcherkez, G. Seed Quality and Carbon Primary Metabolism. Plant Cell Environ. 2019, 42, 2776–2788. [Google Scholar] [CrossRef] [Green Version]
  85. Duermeyer, L.; Khodapanahi, E.; Yan, D.; Krapp, A.; Rothstein, S.J.; Nambara, E. Regulation of Seed Dormancy and Germination by Nitrate. Seed Sci. Res. 2018, 28, 150–157. [Google Scholar] [CrossRef]
  86. Yan, A.; Chen, Z. The Control of Seed Dormancy and Germination by Temperature, Light and Nitrate. Bot. Rev. 2020, 86, 39–75. [Google Scholar] [CrossRef]
  87. Footitt, S.; Huang, Z.; Clay, H.A.; Mead, A.; Finch-Savage, W.E. Temperature, Light and Nitrate Sensing Coordinate Arabidopsis Seed Dormancy Cycling, Resulting in Winter and Summer Annual Phenotypes. Plant J. 2013, 74, 1003–1015. [Google Scholar] [CrossRef] [PubMed]
  88. Finch-Savage, W.E.; Cadman, C.S.C.; Toorop, P.E.; Lynn, J.R.; Hilhorst, H.W.M. Seed Dormancy Release in Arabidopsis Cvi by Dry After-Ripening, Low Temperature, Nitrate and Light Shows Common Quantitative Patterns of Gene Expression Directed by Environmentally Specific Sensing. Plant J. Cell Mol. Biol. 2007, 51, 60–78. [Google Scholar] [CrossRef]
  89. Footitt, S.; Douterelo-Soler, I.; Clay, H.; Finch-Savage, W.E. Dormancy Cycling in Arabidopsis Seeds Is Controlled by Seasonally Distinct Hormone-Signaling Pathways. Proc. Natl. Acad. Sci. USA 2011, 108, 20236–20241. [Google Scholar] [CrossRef] [Green Version]
  90. Yan, D.; Easwaran, V.; Chau, V.; Okamoto, M.; Ierullo, M.; Kimura, M.; Endo, A.; Yano, R.; Pasha, A.; Gong, Y.; et al. NIN-like Protein 8 Is a Master Regulator of Nitrate-Promoted Seed Germination in Arabidopsis. Nat. Commun. 2016, 7, 13179. [Google Scholar] [CrossRef]
  91. Delgado-Sánchez, P.; Ortega-Amaro, M.A.; Jiménez-Bremont, J.F.; Flores, J. Are Fungi Important for Breaking Seed Dormancy in Desert Species? Experimental Evidence in Opuntia streptacantha (Cactaceae): Fungi Break Seed Dormancy in Opuntia. Plant Biol. 2011, 13, 154–159. [Google Scholar] [CrossRef]
  92. Long, R.L.; Gorecki, M.J.; Renton, M.; Scott, J.K.; Colville, L.; Goggin, D.E.; Commander, L.E.; Westcott, D.A.; Cherry, H.; Finch-Savage, W.E. The Ecophysiology of Seed Persistence: A Mechanistic View of the Journey to Germination or Demise: The Ecophysiology of Seed Persistence. Biol. Rev. 2015, 90, 31–59. [Google Scholar] [CrossRef] [PubMed]
  93. Roberts, E.H. Seed Ageing. By D. A. Priestley. Ithaca and London: Cornell University Press (Comstock Publishing Associates) (1986), pp. 304, $41.25. Exp. Agric. 1987, 23, 227. [Google Scholar] [CrossRef]
  94. Zinsmeister, J.; Leprince, O.; Buitink, J. Molecular and Environmental Factors Regulating Seed Longevity. Biochem. J. 2020, 477, 305–323. [Google Scholar] [CrossRef] [PubMed]
  95. Hallam, N.D.; Roberts, B.E.; Osborne, D.J. Embryogenesis and Germination in Rye (Secale Cereale L.): III. Fine Structure and Biochemistry of the Non-Viable Embryo. Planta 1973, 110, 279–290. [Google Scholar] [CrossRef]
  96. Roberts, B.E.; Payne, P.I.; Osborne, D.J. Protein Synthesis and the Viability of Rye Grains. Loss of Activity of Protein-Synthesizing Systems in Vitro Associated with a Loss of Viability. Biochem. J. 1973, 131, 275–286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Benamar, A.; Tallon, C.; Macherel, D. Membrane Integrity and Oxidative Properties of Mitochondria Isolated from Imbibing Pea Seeds after Priming or Accelerated Ageing. Seed Sci. Res. 2003, 13, 35–45. [Google Scholar] [CrossRef]
  98. Bailly, C.; El-Maarouf-Bouteau, H.; Corbineau, F. From Intracellular Signaling Networks to Cell Death: The Dual Role of Reactive Oxygen Species in Seed Physiology. Comptes Rendus Biol. 2008, 331, 806–814. [Google Scholar] [CrossRef] [PubMed]
  99. Corbineau, F.; Gay-Mathieu, C.; Vinel, D.; Côme, D. Decrease in Sunflower (Helianthus annuus) Seed Viability Caused by High Temperature as Related to Energy Metabolism, Membrane Damage and Lipid Composition. Physiol. Plant. 2002, 116, 489–496. [Google Scholar] [CrossRef]
  100. Kibinza, S.; Bazin, J.; Bailly, C.; Farrant, J.M.; Corbineau, F.; El-Maarouf-Bouteau, H. Catalase Is a Key Enzyme in Seed Recovery from Ageing during Priming. Plant Sci. 2011, 181, 309–315. [Google Scholar] [CrossRef]
  101. Waterworth, W.M.; Bray, C.M.; West, C.E. The Importance of Safeguarding Genome Integrity in Germination and Seed Longevity. J. Exp. Bot. 2015, 66, 3549–3558. [Google Scholar] [CrossRef] [Green Version]
  102. El-Maarouf-Bouteau, H.; Mazuy, C.; Corbineau, F.; Bailly, C. DNA Alteration and Programmed Cell Death during Ageing of Sunflower Seed. J. Exp. Bot. 2011, 62, 5003–5011. [Google Scholar] [CrossRef] [Green Version]
  103. Chen, H.; Osuna, D.; Colville, L.; Lorenzo, O.; Graeber, K.; Küster, H.; Leubner-Metzger, G.; Kranner, I. Transcriptome-Wide Mapping of Pea Seed Ageing Reveals a Pivotal Role for Genes Related to Oxidative Stress and Programmed Cell Death. PLoS ONE 2013, 8, e78471. [Google Scholar] [CrossRef] [Green Version]
  104. Wang, Y.; Li, Y.; Xue, H.; Pritchard, H.W.; Wang, X. Reactive Oxygen Species-Provoked Mitochondria-Dependent Cell Death during Ageing of Elm (Ulmus pumila L.) Seeds. Plant J. 2015, 81, 438–452. [Google Scholar] [CrossRef]
  105. Fleming, M.B.; Richards, C.M.; Walters, C. Decline in RNA Integrity of Dry-Stored Soybean Seeds Correlates with Loss of Germination Potential. J. Exp. Bot. 2017, 68, 2219–2230. [Google Scholar] [CrossRef] [Green Version]
  106. Fleming, M.B.; Patterson, E.L.; Reeves, P.A.; Richards, C.M.; Gaines, T.A.; Walters, C. Exploring the Fate of MRNA in Aging Seeds: Protection, Destruction, or Slow Decay? J. Exp. Bot. 2018, 69, 4309–4321. [Google Scholar] [CrossRef]
  107. Kranner, I.; Birtić, S.; Anderson, K.M.; Pritchard, H.W. Glutathione Half-Cell Reduction Potential: A Universal Stress Marker and Modulator of Programmed Cell Death? Free Radic. Biol. Med. 2006, 40, 2155–2165. [Google Scholar] [CrossRef]
  108. Clerkx, E.J.M.; El-Lithy, M.E.; Vierling, E.; Ruys, G.J.; Blankestijn-De Vries, H.; Groot, S.P.C.; Vreugdenhil, D.; Koornneef, M. Analysis of Natural Allelic Variation of Arabidopsis Seed Germination and Seed Longevity Traits between the Accessions Landsberg Erecta and Shakdara, Using a New Recombinant Inbred Line Population. Plant Physiol. 2004, 135, 432–443. [Google Scholar] [CrossRef] [Green Version]
  109. Long, R.L.; Kranner, I.; Panetta, F.D.; Birtic, S.; Adkins, S.W.; Steadman, K.J. Wet-Dry Cycling Extends Seed Persistence by Re-Instating Antioxidant Capacity. Plant Soil 2011, 338, 511–519. [Google Scholar] [CrossRef]
  110. Kibinza, S.; Vinel, D.; Côme, D.; Bailly, C.; Corbineau, F. Sunflower Seed Deterioration as Related to Moisture Content during Ageing, Energy Metabolism and Active Oxygen Species Scavenging. Physiol. Plant. 2006, 128, 496–506. [Google Scholar] [CrossRef]
  111. Châtelain, E.; Satour, P.; Laugier, E.; Ly Vu, B.; Payet, N.; Rey, P.; Montrichard, F. Evidence for Participation of the Methionine Sulfoxide Reductase Repair System in Plant Seed Longevity. Proc. Natl. Acad. Sci. USA 2013, 110, 3633–3638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Ogé, L.; Bourdais, G.; Bove, J.; Collet, B.; Godin, B.; Granier, F.; Boutin, J.-P.; Job, D.; Jullien, M.; Grappin, P. Protein Repair L -Isoaspartyl Methyltransferase1 Is Involved in Both Seed Longevity and Germination Vigor in Arabidopsis. Plant Cell 2008, 20, 3022–3037. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Paparella, S.; Araújo, S.S.; Rossi, G.; Wijayasinghe, M.; Carbonera, D.; Balestrazzi, A. Seed Priming: State of the Art and New Perspectives. Plant Cell Rep. 2015, 34, 1281–1293. [Google Scholar] [CrossRef] [PubMed]
  114. Lutts, S.; Benincasa, P.; Wojtyla, L.; Kubala, S.; Pace, R.; Lechowska, K.; Quinet, M.; Garnczarska, M. Seed Priming: New Comprehensive Approaches for an Old Empirical Technique. In New Challenges in Seed Biology—Basic and Translational Research Driving Seed Technology; Araujo, S., Balestrazzi, A., Eds.; InTech Open Book Series; InTech: Rijeka, Croatia, 2016; pp. 1–46. ISBN 978-953-51-2658-4. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Seed morphology scheme presenting longitudinal section of sunflower fruit as a non-endospermic seed surrounded by the pericarp and Arabidopsis thaliana as an endospermic seed.
Figure 1. Seed morphology scheme presenting longitudinal section of sunflower fruit as a non-endospermic seed surrounded by the pericarp and Arabidopsis thaliana as an endospermic seed.
Biology 11 00168 g001
Figure 2. Seed imbibition curve showing the three characteristic phases with their main biological processes.
Figure 2. Seed imbibition curve showing the three characteristic phases with their main biological processes.
Biology 11 00168 g002
Figure 3. Early cellular events during imbibition as related to water binding forces.
Figure 3. Early cellular events during imbibition as related to water binding forces.
Biology 11 00168 g003
Figure 4. A model for seed cell polarization regulation in the control of dormancy in sunflower [52]. ABA, abscisic acid; PrxII, cell wall peroxidase III; POX, cytosolic peroxidase; ET, ethylene; Vm, PM potential.
Figure 4. A model for seed cell polarization regulation in the control of dormancy in sunflower [52]. ABA, abscisic acid; PrxII, cell wall peroxidase III; POX, cytosolic peroxidase; ET, ethylene; Vm, PM potential.
Biology 11 00168 g004
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

El-Maarouf-Bouteau, H. The Seed and the Metabolism Regulation. Biology 2022, 11, 168. https://0-doi-org.brum.beds.ac.uk/10.3390/biology11020168

AMA Style

El-Maarouf-Bouteau H. The Seed and the Metabolism Regulation. Biology. 2022; 11(2):168. https://0-doi-org.brum.beds.ac.uk/10.3390/biology11020168

Chicago/Turabian Style

El-Maarouf-Bouteau, Hayat. 2022. "The Seed and the Metabolism Regulation" Biology 11, no. 2: 168. https://0-doi-org.brum.beds.ac.uk/10.3390/biology11020168

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