Open access peer-reviewed chapter

Responses of Neotropical Savannah Plant Species to Abiotic Stresses: A Structural and Functional Overview

Written By

Silvana Aparecida Barbosa de Castro and Vinícius Coelho Kuster

Submitted: 27 May 2020 Reviewed: 04 September 2020 Published: 08 October 2020

DOI: 10.5772/intechopen.93891

From the Edited Volume

Abiotic Stress in Plants

Edited by Shah Fahad, Shah Saud, Yajun Chen, Chao Wu and Depeng Wang

Chapter metrics overview

503 Chapter Downloads

View Full Metrics

Abstract

Plants under field conditions are subject to different types of abiotic stresses such as drought, salinity, and light excess that adversely affect their growth and survival. In addition, several studies have pointed out the effect of climate change such as an increase in the concentration of atmospheric CO2, as well as an increase in global temperature on the distribution and wealth of plants. Adaptation to abiotic stress and survival occurs on different scales, at the cellular level for each individual, and requires a range of strategies, whether morphological, physiological, molecular or structural. Such strategies may be determinant in the distribution of plant species in natural habitats, depending on ecological adaptations shaped by the evolutionary history of species. In this chapter, we discuss recent information about mechanisms of plant adaptation to abiotic stress in the Neotropical savannah based on the cell and individual scales.

Keywords

  • ecophysiology
  • water stress
  • thermotolerance
  • high luminosity
  • soil nutrients
  • fire
  • Cerrado

1. Introduction

Plants growing under natural conditions are permanently subject to different types of environmental stresses such as drought, nutritional deficiency, salinity, heat stress and, more recently, the anthropogenic pollutants of ecosystems that negatively affect plant growth and survival [1, 2, 3, 4, 5]. Moreover, studies have pointed out the effect of climate changes such as the increase of atmospheric CO2 concentration and global temperature (2 to 4°C) on the distribution and richness of plant species [6, 7, 8]. The survival of plant species under such conditions is constantly threatened, leading them to use different cellular and whole plant mechanisms in order to minimize damage and adjust growth to adverse environmental conditions [9, 10]. Plant responses to abiotic conditions have been widely investigated in different ecosystems worldwide, with this issue gaining great prominence for savannas due to the high complexity of abiotic factors that affect the native species of these systems. Among the world’s savannas, the Brazilian one, also called Neotropical savannah or Cerrado sensu lato, is one of the most diverse in the world in terms of vegetation and plant species types, recognized as one of the biodiversity hotspots in the world [11, 12]. The Neotropical savannah flora comprises some 12,000 plant species, many of them endemic, two-thirds of which are herbaceous species or small shrubs [13]. The Neotropical savannah covers approximately 22% of the Brazilian territory (Figure 1) comprising several physiognomies from open grasslands to woodlands in a mosaic determined by abiotic factors such as edaphic and fire ones [14, 15]. The Campo limpo (grasslands) and Cerradão (tall woodland) are both extreme physiognomies, with great predominance of herbaceous and arboreous strata, respectively [14]. The Campo sujo (shrub savannah), Campo cerrado (wooded savannah), and Cerrado sensu stricto (short woodland) occur between them, with a gradual growth of biomass and larger plant species strata [14]. A mountain physiognomy, called rupestrian fields, also occurs and is restricted to areas of more than 900 m of altitude [16].

Figure 1.

Map of Brazil showing the occurrence of the Neotropical savannah, which covers approximately 22% of its surface area (Adapted from WWF Brazil, 2020).

The diversity of life-form strategies of the Neotropical savannah plant species is associated with the response to many abiotic factors, especially the climatic seasonality with a well-defined dry and rainy seasons (Figure 2). The rainy season normally lasts from October to March, and the dry season from April to September, with an annual rainfall average of about 1,477 mm, a minimum temperature of 16°C and a maximum temperature of 26°C over the past thirty years (Figure 2) [17, 18]. High temperatures and irradiance, oligotrophic soils and high content of solubilized Al3+ are environmental features of the Neotropical savannah, which select local plant species and induce them to develop physiological and morphological adaptations to this harsh environment [19, 20]. Adaptation to and survival of abiotic stress occur on different scales from the cellular to the individual levels and require a range of strategies, whether morphological, physiological, molecular or structural, and can be determinant of species distribution in the natural habitats, depending on ecological adaptations shaped by the evolutionary history of the species.

Figure 2.

Rainfall (mm) and temperature (°C) data of the Neotropical savannah during the rainy season (represented in blue area) and the dry season (represented in yellow area), over the past thirty years. Ps.: Black circles indicate the maximum temperature and white circles the minimum temperature (data obtained from Instituto Nacional de Meteorologia - INMET, 2020).

Based on recent studies on Neotropical savannah flora, the focus of this review is on the dynamics and processes related to plant responses to the most prominent abiotic environmental conditions in this ecosystem, mainly highlighting water, temperature, light, nutrient deficiency and fire approaches. We report here information about the different adaptation mechanisms of plant species under the abiotic stress of the Neotropical savannah, from the cell to the individual scales.

Advertisement

2. The water stress seasonality changes the response of plant species during the seasons and along the day

Drought is considered to be one of the major abiotic stresses since it influences the distribution of plant species and limits the productivity of the ecosystems [4]. All aspects of plant development, i.e. germination, cell division, cell elongation, plant growth, and metabolic activities, are affected by a reduction in soil water content, which subsequently affects the physiological, biochemical, morphological and molecular processes of plants [21, 22]. Water flow from the soil to the roots depends of several factors, such as, potential water gradient between the soil and the plant, the hydraulic conductivity of the soil, the soil moisture content, and the atmospheric demand, that together with leaf transpiration create a tension in the xylem contributing to the creation of this water potential gradient [22, 23].

Regarding water stress, physiological and biochemical responses are more immediate in order to avoid cell dehydration, to maximize the efficiency of water use or to tolerate dehydration [1, 24]. The ability to maintain higher values of cellular water content is known as a strategy of plants sensitive to drought conditions. On the other hand, the ability to survive with lower values of cellular water content is called drought tolerance and requires the expression of molecular mechanisms since the initial growth of the seedling [25, 26]. Drought-tolerant plant species can be found in ecological niches with water seasonality, especially in the Neotropical savannah [27]. Drought-sensitive plant species also occur at the same sites as those of drought-tolerant ones; however, they use different molecular, biochemical, and physiological mechanisms to prevent desiccation, and thus deal with the lack of water in the soil. Drought-tolerant plants are able to limit damage to a repairable level in order to maintain physiological integrity due to the occurrence of repair mechanisms triggered by the dehydration/rehydration cycle [27, 28]. During long periods of drought, maintaining cell homeostasis and withstanding dehydration seems impractical for plants, directly impacting their growth rates [4, 27].

The Neotropical savannah abiotic conditions such as high irradiance, elevated air temperature, and low relative humidity impose a high evaporative demand on plants during the dry season while at the same time water is exhausted in the upper layers of the soil [29, 30]. Water use by Neotropical savannah woody plants is limited by a combination of physiological, plant architectural and tree density constraints evidencing that water use by the plant community seems to be more complex than in other savannas [30]. Even in the dry season, the water supply remains constant in the deeper soil layers, allowing access to water for tree species with long roots (in trees and shrubs) and differences between functional groups may be more related to the topography of an environmental continuum in which grass species are most subject to water restriction (Figure 3) [30, 31]. The hydrological niche segregation, defined as a fine scale partitioning of space for soil moisture gradients or different strategies of water acquisition, has been pointed out in different plant communities [32, 33] including the Neotropical savannah physiognomy [34, 35]. A marked decrease in water potential during the dry season has been reported for several species of the Neotropical savannah both in predawn and midday measurements [31, 35, 36, 37]. However, the water potential can vary according to the microhabitat, with plant species in mesic microhabitats showing smaller drops in the values of water potential even during the dry season (Figure 3). Costa et al. [38] also observed a decrease in water potential in Hymenaea stigonocarpa (Fabaceae) subjected to water stress, followed by recovery after rehydration. In an experimental study with Laivosiera campos-portoana (Melastomataceae), França et al. [39] reported that drought stress triggered a decrease of stomatal conductance and that osmotic adjustment played an important role in keeping high relative water content (RWC) values. These data suggest that the maintenance of high RWC values in plant species, despite a decrease in xylem water potential during the dry season, may be related to adjustment to stress, allowing survival in different microhabitats.

Figure 3.

Schematic presentation of the main water parameters of tree and grass species from the Neotropical savannah during the rainy (represented in blue area) and dry (represented in yellow area) seasons. Trees are able to capture water from deep layers of soil in the dry season, but there is still a certain water restriction that may affect water potential and stomatal conductance in some species. Grasses have lower water potential values and greater restriction of stomatal opening than tree species in the dry season. Brown arrows indicate an increase in rates while the red arrows show a decrease. Abbreviations: ψ pd - Predawn leaf water potential; GS - Stomatal conductance; VPD - Relative vapor pressure deficit.

Plant responses to water deficit occur in different forms and scales, e.g. morphological, physiological, biochemical, and molecular ones, which may occur at the same time or at different times. For instance, the synthesis of cellulose microfibrils, hemicelluloses, and pectins in the cell wall may be changed under water stress, especially regarding the cell elongation pattern [40, 41, 42]. Strategies for adaptation and survival under water deficit conditions are also associated with the production and accumulation of organic solutes such as glycine-betaine, proline, and soluble sugars, which protect the integrity of the membrane and promote osmotic adjustment [43, 44, 45, 46, 47]. In addition, an increased content of several amino acids such as glutamic acid, glutamine, lysine, proline, serine, tryptophan, tyrosine and valine has been reported after water stress imposition in Neotropical savannah species [47, 48]. Proline is an important osmoregulator that protects plants submitted to abiotic stresses, favoring osmotic adjustment and an increase in the other osmolytes [20]. Souza et al. [48] detected increased carbohydrate content in Hymenae courbaril (Fabaceae) roots subjected to water deficit, which appeared to be directly linked to the osmotic adjustment of the cell. Some studies have indicated an accumulation of sucrose, raffinose and arabinose during the dry season, while glucose and fructose levels decreased with the reduction of water availability in the soil, suggesting that these are used for sucrose synthesis [49, 50].

The plant antioxidant system also plays a key role in water stress by responding with the production of enzymes such as superoxide dismutase (SOD), catalase (CAT), ascorbate peroxidase (APX), and glutathione peroxidase (GPX), as well as non-enzymatic components such as ascorbate, glutathione, tocopherols (vitamin E) and carotenoid pigments [51, 52]. Imbalance in the production of reactive oxygen species (ROS) has been demonstrated to affect negatively cell membrane stability, with oxidative damage to carbohydrates, lipids, amino acids, proteins, and nucleic acids [51]. ROS play a key role in the process of plant acclimation to abiotic stress such as signal transduction, which regulates different pathways during plant acclimation to stress [53]. Moreover, many plants may rapidly produce and scavenge different forms of ROS, inducing fast and dynamic changes in ROS level [51]. Under normal growth conditions, ROS are continuously produced and scavenged in organelles such as chloroplasts, mitochondria, peroxisomes and apoplasts [54]. However, the balance between ROS-producing pathways and ROS-scavenging mechanisms may be disrupted when plants experience environmental stress such as drought [52, 55, 56].

Regarding the photosynthetic apparatus, Assaha et al. [57] reported a super excitation of photosynthetic pigments after ROS accumulation in plants under severe water deficit. For instance, increase in oxidative stress (i.e. increase of H2O2) and in SOD, CAT, and APX activity in roots was reported in two Neotropical savannah tree species subjected to water stress imposed by competitive species [58]. As well, Vieira et al. [58] also showed an increase in the activity of enzymes of the antioxidant system in Vatairea macrocarpa (Fabaceae) subjected to water stress. Besides, ROS detoxification systems, other strategies are essential for plant tolerance to drought, such as a reduction in chlorophyll synthesis and in the assimilatory apparatus components, as well as a balanced sugar pentose-phosphate pathway [59, 60, 61].

The seasonality of water availability can also result in dynamic photoinhibition in woody plants [62], but not in herbaceous species [63, 64, 65]. A reduction of stomatal conductance and of net CO2 assimilation has also been reported in Barbacenia Purpurea (Velloziaceae) submitted to water stress, but recovery was detected hours after rehydration [50]. Decreased stomatal conductance followed by a reduction in photosynthetic rates in plants under moderate water deficit has been reported [66]. Midday depression is related to excessive sunlight exposure and has been shown for a wide range of plant species of the Neotropical savannah, with impact on photosynthesis and on CO2 assimilation [62, 67, 68]. The reduction of the Fv/Fm ratio, of maximum variable fluorescence (Fv) and of maximum fluorescence (Fm) during the dry season seems to be a consequence of a combination of strong stomatal control, especially in shrub species, high irradiance, and low soil moisture conditions [35, 67]. Stomatal control may prevent the dissipation of excessive energy and lead to chronic photoinhibition in some Neotropical savannah plant species [68]. Chronic seasonal photoinhibition has been shown for some plant species in the xeric microhabitat of the Neotropical savannah, as reported for Melastomataceae species, although it does not occur in mesic habitats [35, 63, 69]. These data reveal the importance of the specific characteristics of the microhabitat in the physiological studies of plants despite the seasonal variation of water availability. Several factors such as microhabitat, root system, functional group and topography directly influence water use strategies by the plant community, causing the Neotropical savannah to be more complex than other savannas.

Advertisement

3. Thermal stress triggers morphological and physiological damage and acts as selecting factor

Plants have different temperature domains that initially determine their geographic distribution [70] but with the progressive increase in air temperature, the temperature domains become displaced to colder regions, forcing plant species to move away from their current distribution [71, 72]. The increase in temperature may cause several types of damage to plants in different stages of life, from germination to adulthood [73, 74]. The temperature range considered optimal for photosynthesis in C3 plants is between 18°C and 30°C, while temperatures above 30°C are considered to be heating zones [70, 75]. Thermotolerance acquisition occurs when a plant is exposed to low or high temperatures for a short period of time, up to a limit where no fatal injuries occur [8] being thermotolerance fundamental for the plant to deal with greater variation in temperature under certain conditions.

Morphological damage has been reported for vascular plants as a response to heat stress, with emphasis on leaf and branch burn, foliar senescence and abscission, inhibition of shoot and root growth, discoloration, and fruit damage [76]. Associated with the damage triggered by heat stress, the response to rising temperatures varies widely among plant species, functional groups and life strategies [77, 78]. For instance, fast-growing plants from tropical forests in high light environments tend to be more tolerant of high temperatures than slow-growing plant species typical of shaded places in the understory [77]. According to Franco et al. [79], there is controversy over the limiting processes that control photosynthesis at high temperatures, with different responses in the metabolism of C3 and C4 plants. In particular, under high temperatures, the photosynthetic rate (A) can increase up to a point above which it starts to be inhibited and can reach zero [77]. At higher temperatures, plants show elevated stomatal conductance as a way to dissipate excess heat on leaf surfaces. However, in combination with water stress, greater control of stomatal opening can limit heat dissipation with an impact on photosynthesis [60, 80, 81]. The reduction of photosynthesis occurs due to damage to the electron transport chain associated with increased fluidity of the thylakoid membrane and damage to photosystem II (PSII) [82, 83]. PSII is highly sensitive to temperature and damage can be observed after a few minutes to a few hours of exposure to heat [60, 82, 84, 85]. A reduction in the membrane lipid content and an increase in lipid degradation during the late period of stress with formation of ROS exposure has been reported and the balance between ROS production and elimination is rapidly disturbed, leading to an increase in ROS content [86]. Changes in chlorophyll content in plants exposed to high temperatures have also been reported, possibly due to inhibition of chlorophyll biosynthesis for the destruction of numerous enzymes involved in their biosynthesis [87].

A highly consolidated mechanism of heat acclimation and rapid response to high temperature in plants is the production of heat shock proteins (HSPs), which are key for thermotolerance acquisition [88]. HSPs protect intracellular proteins from denaturation and preserve their stability, thus acting as chaperones [88, 89, 90]. According to Ohama et al. [90], HSPs include HSP100, HSP90, HSP70, HSP60 and small HSPs that play a key role in protein quality regulation by renaturing a variety of proteins denatured by heat stress. Together, HSPs and ROS scavenging enzymes are the major functional proteins induced by heat stress [90].

Duarte et al. [91] detected increased oxidative stress at high temperatures in seedlings of the Neotropical savannah plant species Vriesea friburgensis and Alcantarea imperialis (Bromeliaceae) due to the accumulation of H2O2, which in turn led to a decrease in the number of leaves and in biomass and to a high mortality rate in Vriesea friburgensis when submitted to treatment at 35°C. Under field conditions, Neotropical savannah plant species show photoinhibition during the hottest hours of the day, especially in combination with seasonal water stress [62, 92]. Changes in morphological traits have been observed for other Neotropical savannah plant species, such as germination in Dipteryx alata (Fabaceae) in an experiment involving increasing temperatures [93] and in the response to zinc absorption in Pterogyne nitens (Fabaceae) [94]. Changes more directly related to photosynthesis, water use efficiency, growth rate, factors that can alter the competitive capacity of species, have been observed [73, 95, 96].

Model-based studies predict changes in plant species distribution and loss in areas with greater biodiversity due to increased worldwide global temperature [97, 98, 99, 100], also affecting the Neotropical savannah [100, 101]. Chaves et al. [102], in a study of a restricted plant species from the rupestrian field, showed that although some plants were no longer sensitive to the increase in temperature, they showed lower thermal tolerance and less plasticity compared to plant species of wide distribution, with a possible impact of global climate changes on this species. A recent modeling study evaluating the effect of habitat loss and climate change on 2,354 species of restricted distribution (including the Neotropical savannah) showed that 70 to 85% of them are at high risk of extinction [103]. With the increase in global temperature it is possible that plant species may be following three paths: adapting through phenotypic plasticity, migrating to higher altitude, or becoming extinct [7, 101]. Bueno et al. [104] showed that the greatest expansion of the Neotropical savannah occurred precisely during periods of higher temperatures (Last Interglacial). However, this result was certainly based on climate stability that allowed a large number of plant species to persist over time to the detriment of areas with major climatic changes where species might not have had enough time to adapt or to move to regions with more favorable conditions [105, 106].

Multiple environmental factors vary with increasing altitude, with temperature being perhaps the most important one because it directly influences biochemical processes related to photosynthesis and water status. Therefore, studies that evaluate the physiological responses of plant species at high temperatures are important because they provide data that can generate models of future plant species distribution considering different scenarios of temperature change, and because they serve as the basis for public policies aimed at protecting areas with plant species most vulnerable to global climate change.

Advertisement

4. Neotropical savannah plant species must deal with spatial and temporal light heterogeneity

Light is one of the abiotic factors that influence the survival, development and growth of plants, and several traits are highly plastic in the presence of substantial variations in the amount of light received, especially traits related to light capture and photosynthesis [107, 108, 109, 110]. Light also shows spatial and temporal heterogeneity as an important factor underlying niche partitioning among plants [111, 112], which has created plants adapted to a variety of light intensity, such as sun-adapted and shade-tolerant plant species [113]. For photosynthesis, the flux of photons in the 400–700 nm range is the most relevant variable to be considered in studies that evaluate changes in light intensity on the development and growth of plants [110]. Under high-light conditions, plants possess a variety of mechanisms that maintain the operational photosynthetic system, such as carotenoids, violaxanthin, antheraxanthin and zeaxanthin, which play an important role in the dissipation of excitation energy at times when light levels exceed the plant capacity for photosynthetic electron transport and carbon fixation [114, 115]. However, plants exposed to high radiation may suffer increased leaf temperature, resulting in photoinhibition of Photosystem II (PSII) [116, 117].

In tropical regions, the irradiance levels normally exceed those that plants use in photosynthesis, so that many plants have efficient mechanisms preventing or minimizing the damage to the photosynthetic apparatus caused by high light availability [118]. Nevertheless, different light regimes in each physiognomy can promote the establishment of different plant species or functional traits [119]. Regarding the Neotropical savannah, most plant species are well adapted to high light intensity, and at the same time they may have low plasticity for adaptation to and growth in shaded environments [120, 121]. Under contrasting light conditions, some traits such as stomatal conductance, chlorophyll and carotenoid contents have been shown to have higher rates of plasticity in plant species growth and in different seasons (i.e. dry and rainy seasons) [120, 122]. According to Franco et al. [92], a balance between the use of light absorbed for photosynthesis and the dissipation of potentially harmful excess light is essential for Neotropical savannah plant species since they still have to deal with other abiotic factors (i.e., high air temperature, low relative humidity and seasonal drought). Regarding the photosynthesis apparatus, Franco et al. [92] also reported that woody plant species are capable of maintaining the functionality of PSII under conditions of high irradiance and high evaporative demand of the atmosphere during long sunny periods at the end of the dry season. In contrast, in a study of Hymenaea stigonocarpa (Fabaceae) under high light intensity, Fv/Fm values decreased to 0.64, indicating the occurrence of photoinhibition as well as a lower water potential value [123]. In a study of Copaiba langsdorffii (Fabaceae) growing under full sun and shade conditions, Ronquim et al. [119] observed that juvenile plants under full sun conditions showed higher maximum net photosynthetic rates, light compensation and saturation points than juvenile plants growing under shade conditions. Other leaf traits associated with light have been pointed out, such as low specific leaf area (SLA) values and high carotenoid and chlorophyll content in plant species from open areas, which may represent adaptive mechanisms to high luminosity environments [36, 79, 121].

In another approach, recent studies have considered the impact of plant invasions in open areas of the Neotropical savannah caused by the widespread expansion of forest in the absence of fire, which tend to change the availability of light for native species already adapted to high radiation [121]. In areas with invasion and expansion of forest, changes in leaf area index (LAI) have been reported, as well as changes in the photosynthetically active radiation available to Neotropical savannah plant species [124]. Moreover, Rossatto et al. [121] showed that invasion of originally open savannah areas by forest trees has an impact on the persistence of savannah plant species, which need greater availability of light to maintain a positive carbon balance. Under these conditions, it is expected that grasses and herbaceous species, already adapted to high light availability, will not be able to maintain the CO2 assimilation necessary for the maintenance of root carbon stock, which is essential for adaptation to fire and seasonal water stress.

Advertisement

5. Neotropical savannah plants live in nutrient-poor soils with aluminum phytotoxicity

The Neotropical savannah soils are considered to be among the most chemically infertile soils in the world [124, 125], being generally dystrophic, acidic and poor in available nutrients [126, 127], as well as having high levels of exchangeable aluminum (Al) [127, 128]. Soils are mainly represented by well-drained oxisols [129], named “latossolos” in the Brazilian soil classification system [130], followed by sandy entisol, ultisol and cambisol soil types [129, 131]. Many relevant plant nutrients such as K, Ca, Mg, N, S, and P, are deficient in the soil of most Neotropical savannah physiognomies due to the nutritional deficiency of the initial matrix rocks and the intense weathering that leads to daily loss of nutrients [127]. Although Neotropical savannah soils are generally defined as dystrophic, mesotrophic physiognomies have also been reported [132, 133], influencing the variation of floristic structure, as reported by Neri et al. [133] when differentiating dystrophic and mesotrophic Cerradão physiognomies.

In the nutritional approach to studies of the Neotropical savannah, aluminum phytotoxicity and the responses of native plants to low nutrient availability, mainly nitrogen and phosphorus, are some of the most highlighted topics, which will be discussed here. Al is one of the main metals present in the crust of the Earth [134], being a limiting factor of plant development [135, 136]. Most Al is available in soils in the form of non-phytotoxic elements such as Al oxides and Al silicates [137]. However, under acid conditions (pH < 5), Al is solubilized to Al3+ [138, 139, 140], as well as to Al(OH)2+, Al(OH)2+ and Al(OH)63+ [141], all known as phytotoxic forms. The relationship between Al phytotoxicity and acidic soils stands out mainly in the Neotropical savannah, where some physiognomies such as Neotropical savannah sensu stricto and associated open physiognomies contain acid soils [126]. High Al concentrations, especially Al3+, occur in dystrophic soils, while soils almost free of exchangeable Al occur in mesotrophic physiognomies with pH > 6.0 [127, 133, 142].

Al3+ is extremely toxic for the majority of cultivated plants, but, as expected, the native Neotropical savannah plants are highly tolerant to it [133] and may be divided into Al-accumulating and non-accumulating plants [143, 144, 145]. Plants that store 1000 mg kg−1 Al are defined as Al-accumulating species [127, 144, 146], which can accumulate Al in roots or aerial parts [142]. For instance, Qualea grandiflora (Vochysiaceae), Qualea cordata (Vochysiaceae) and Miconia albicans (Melastomataceae) have been reported to be Al-accumulating species [132, 147]. Moreover, concentrations of about 20,000 mg Al kg−1 dry leaves have been observed for some native Al-accumulating plant species of the Neotropical savannah [132, 143] such as noted for Vochysia tucanorum (Vochysiaceae) [132]. Many Al-accumulating plant species not only tolerate Al but also need it for their growth and development, as is the case for some plants of the Rubiaceae, Melastomataceae, and Vochysiaceae families [148]. Thus, Al-accumulating plant species, in many cases, reduce their growth rates when subjected to the absence or a low concentration of aluminum [127, 147, 149]. On the other hand, many plant species do not accumulate more than 1000 mg/kg−1 Al in their leaves, as reported for 24 out of 32 plants from the central part of the Neotropical savannah, such as Caryocar brasiliense (Caryocaraceae), Vellozia squamata (Velloziaceae) and Roupala montana (Proteaceae) [143].

Reduced developmental rates and physiological changes may be some symptoms related to high Al concentrations in non-Al accumulating Neotropical savannah plants. The inhibition of root growth [103], the lack of lateral roots [150], and a low biomass of stems and leaves [141, 148, 151] have been demonstrated in plants growing in the presence of high Al concentrations. According to Čiamporová [152], the inhibition of root growth may be due to Al impregnation of root cell walls that promotes unequal expansion and changes in tissue organization. This process occurs in many Al-sensitive plants, while in Al-tolerant species the high Al deposition on the cell wall normally does not inhibit root elongation, as reported for soybean seedlings [153]. Organic acid exudation from the roots has been cited as an important strategy for Al exclusion [150, 154], which avoids Al uptake and the associated structural and physiological symptoms. Bittencourt et al. [155] have shown that Styrax camporum (Styracaceae) releases organic acids in the roots as a defense mechanism against high Al accumulation. However, this process only works when the plant is exposed to low to moderate Al levels.

Regarding the physiological responses, previous studies have demonstrated that CO2 assimilation rate and stomatal conductance were reduced in Al non-accumulating species submitted to high Al concentrations [150, 151, 156]. The leaf gas exchange performance of Al-sensitive species has been associated with the lack of lateral roots [150, 157, 158], which are the plant’s water uptake site. The lack of a water supply from the roots reduces its amount in the mesophyll, leading to low gas exchange [159]. Associated with reduced gas exchange performance, the CO2 assimilation rate was also reduced in some plant species submitted to high Al doses [160], as reported for Citrus [161, 162] and for two sorghum cultivars [163]. From a photochemical perspective, quantum yield, maximum variable fluorescence ratio (Fv), minimum fluorescence (Fo), and Fv/Fm have been reduced in the presence of high rates of Al [164], an event that seems to be related to the impairment of PSII photochemistry [163] to the reduction of PSII electron transport [165], and to other changes in photochemical processes. However, substantial changes in Fv/Fm values have not been observed in Al accumulating and Al non-accumulating species subjected to high Al concentrations [147, 150].

In contrast to Al, nitrogen (N) and phosphorus (P) are reported as the limiting factors in the Neotropical savannah [166] both due to the old age of the soil and to the recurring effect of fire, which contributes to the volatility of N parts [167, 168]. N is a prevalent element in living systems and is a constituent of phytohormones, amino acids, proteins, nucleic acids, and enzymes [169]. It acts on the initial growth processes, helping with the synthesis of deoxyribonucleic acids and the replication of chromosomes [169]. P is also required for plant development since it is involved in metabolic regulation, such as amino acid synthesis and protein activation [170], and is also a constituent of ATP/ADP and phospholipids [171]. Thus, N and P are nutrients required in large amounts for initial plant development [172]. According to Reis et al. [173], both nutrients are normally present in limited amounts in Neotropical savannah soils since low N availability may be related to the presence of minimal organic matter and to the fact that P is low in soils containing iron and aluminum oxides [174], both typical conditions of most Neotropical savannah soils. N and P are constituents of photosynthetic process-related proteins such as plant pigments [141], thus representing fundamental elements for the maintenance of photosynthetic quantum yield and biochemical steps. For instance, Eugenia dysenterica (Myrtaceae) seedlings showed increased photosynthetic activity and growth with N fertilization at rates between 50 and 200 mg dm−3, while P fertilization at rates of 100 mg dm−3 and 200 mg dm−3 increased growth and photosynthetic activity, respectively [173]. In addition, some findings suggest that the growth of many plant species is reduced in Savannah soils, although the chemical attributes per se do not seem to exclude the occurrence of plants in Neotropical savannah physiognomies [175]. Despite the low availability of N and P, plant species growing in Neotropical savannah soils have a variety of mechanisms dealing with shortages of these nutrients. For instance, P is allocated to photosynthetically active mesophyll cells, showing higher photosynthetic P-use efficiency (PPUE), association with mycorrhizal species, different ways of nutrient uptake, and an efficient use of nitrogen [168, 176].

From another perspective, over the last decades, the Neotropical savannah has suffered enormous losses of its original cover (more than 50%) [11, 177]. Moreover, the change in land use, such as the introduction of extensive pecuary and immense extension of soybean crops, has contributed to the growing supply of these nutrients in the Neotropical savannah soils [178]. Studies have shown that the increase in N and P favors the increase in biomass, the decrease in species richness, the establishment of invasive species and changes in plant structure (root/shoot ratio, and root structure) and in competitive relationships [179, 180, 181]. It has also been shown that ancient ecosystems with less P availability in the soil are the richest in plant species [182, 183]. Thus, it is expected that an increase in P supply in the Neotropical savannah soils may have an impact on plant species richness in the medium and long-term.

Advertisement

6. Fire as a selective agent for plant communities

Fire has shaped different terrestrial environments since early geological ages and has been part of Earth’s system since the origin of land plants in the Silurian period [184, 185]. In the Neotropical savannah, fire is one of the determining factors that select and structure plant communities and that promote the cycling of nutrients, representing a fundamental abiotic agent for the maintenance of open ecosystems [186, 187]. In the phytophysiognomies of the Neotropical savannah, fire normally occurs at the end of the dry season [188] as a result of the highly combustible material of dry herbaceous plant organs (Figure 4), while during the rainy season, natural fires can be caused by lightning [187, 189]. Although fire is a natural and periodic component of this environment, anthropic activities have intensified its action and made it more frequent [190], leading to cyclic burning of the herbaceous layer and a marked reduction of the maintenance and renewal of trees and shrubs, progressively decreasing their density [191]. Thus, the main problem with fire is that it is seen as a bad event when its regime (frequency, extension, and intensity) is somehow changed [187, 192].

Figure 4.

Overview of the Neotropical savannah plant species in the pre-fire (blue circle) and post-fire (pink circle) periods, which coincided with the rainy and dry seasons, respectively. At a pre-fire time during the rainy season, the herbaceous stratum exhibits dry and flammable aerial organs, while tree and shrub species have green and photosynthesizes leaves. At a post-fire time during the dry season, the leaves of the herbaceous stratum regrow, while many tree species are still leafless due to leaf loss caused by heat, and with a stem blackened by the action of fire.

Given the seasonality of fire in the Neotropical savannah, plants possess physiological, morphological, ecological, chemical, and phenological adaptations to deal with it. These adaptations, as well as adaptation to drought (see [193]), can enable plant species to escape from fire, as is the case for plants with a short life cycle, or to resist fire, as is the case for tree species with extra protective bark [194]. As an escape strategy, many herbaceous species synchronize seed production before annual burning periods and enter a type of dormancy that only ends after high-temperature exposure [192], as reported for different plant species. For instance, the dormancy of Sporobolus cubensis (Poaceae), a herbaceous species of the Neotropical savannah, was broken by exposure to 200°C, indicating the need of fire temperature for its germination [195]. However, many seeds and underground plant organs may be protected against scorching since the heat from fire events does not reach deeper layers of the soil [196]. In general, a fire event is followed by a high germination rate and seedling development since the supply of nutrients increases from the ashes and more light reaches the plant from the opening of canopies [192, 197].

In a phenological approach, fire can affect vegetative and reproductive phenophases [186], leading to leaf abscission by the heat created from the fire, as well as decreasing the numbers of flowers and fruits [191, 198]. However, Silvério et al. [199] reported that the negative effects of fire on the vegetative phenophases were normally followed by an increase in sprouting and the production of leaves, revealing an epigeous recovery mechanism from fire damage. In a reproductive approach, Pilon et al. [200] reported a similar pattern, with fire triggering the production of flowers and fruits in 79% of 47 plant species such as grasses, forbs, and subshrubs. Moreover, several studies have reported flowering in several plant species after the passage of fire [201, 202] and, more recently, Fidelis et al. [203] reported massive flowering in Bulbostylis paradoxa (Cyperaceae) within just 24 h after the fire. Flowering after fire certainly gives and advantage, since it allows the plant to use available resources without germination barriers and in new and fresh open areas for its establishment [204].

Many structural plant features have been proposed as important tools for fire event survival. Underground roots and stems with carbohydrate, nutrient and water storage have been reported for herbaceous and sub-shrub Neotropical savannah species, allowing them to regrow after the fire has passed [205]. For instance, diffuse underground systems, rhizophores, and xylopodium have been reported for 7 Asteraceae species from the Neotropical savannah [206], which can support the individual life during the dry season and after fire events. Xilopodium, as well as other underground systems, have been found in many species as reported for the Brosimum gaudichaudii (Moraceae) shrub [207], the Ocimum nudicaule herb (Lamiaceae) [208] and three Erythroxylum species (Erythroxylaceae) [209], and may indicate be a convergent trait of many plant species in this environment, often related to species dominance in physiognomies.

Regrowth after fire events requires carbon [79] for the creation of new tissues and organs; however, burning that reaches the leaves reduces the photosynthetic area of the plants [210], consequently decreasing the capture of atmospheric CO2. The loss of leaf area can be total in herbaceous and shrub plants and partial in some sub-arboreal plants during fire events, whereas tree branches normally are not burnt but are damaged by the hot airflow [79]. In this respect, underground reserve organs may be a fundamental carbon source, mostly consisting of carbohydrates such as starch, fructan, glucose and sucrose [189, 206]. Complex carbohydrates such as starch are hydrolyzed to soluble sugars by hydrolytic enzymes such as the exoamylase β-amylase, which cleaves α-1,4 linkages, and the debranching enzyme isoamylase, which cleaves α-1,6 linkages [211] and the product used for plant development. After leaf regrowth, the new leaves of some plant species may return with higher photosynthetic rates, stomatal conductance and nitrogen concentration than during the pre-burn period [212], with high vigor and potential for growth [213]. However, these responses did not occur in Vochysia cinnamomea (Vochysiaceae), a native species of Neotropical savannah, since fire did not exert a direct influence on its physiological processes [214].

The bark structure and thickness of tree plants has been cited for years as one of the greatest protective traits against fire in the Neotropical savannah, preventing damage to the vascular cambium [215] and deformations of the xylem [216] and phloem. The thick corky bark may protect internal stem tissues [217], reducing mortality and maintaining water and solute transport during and after the fire episodes [218]. On this basis, the bark properties have been studied across 31 species from the Neotropical savannah, providing shreds of evidence that the outer bark acts as mechanical support and defense against pathogen attack and the inner bark as a place related to storage and transport of water [218]. Other less frequent and more specific characteristics have also been discussed regarding fire survival. For instance, Vellozia variabilis (Velloziaceae) has resin glands on its leaves that produce flammable substances capable of starting quick-burning on the more protected plant base, preventing damage to more sensitive upper plant portions [219]. Due to these plant specificities, as well as abiotic and biotic conditions, the Neotropical savannah is a potential environment for new structural and ecological discoveries.

Finally, the occurrence of invasive tree species in open areas of the Neotropical savannah has altered the microclimate conditions, affecting grasses and shrubs and consequently the dynamics of fire occurrence in the Neotropical savannah [124, 220]. This cascade of events will certainly have a negative impact on the plant and animal communities of the Neotropical savannah, changing the local landscape.

Advertisement

7. Conclusions

In the current chapter, we report the most common types of abiotic stress in Neotropical savannah plant species, which can simultaneously influence and determine the different functional types of plants. Plant species are affected and respond differently to abiotic stress in the Neotropical savannah, mostly in terms of more specific characteristics such as microhabitat, seasonality, topography and anthropic interference. Moreover, Neotropical savannah plant species have evolved over thousands of years subject to various types of abiotic stress (i.e. water seasonality, high temperature and light availability, soil with a deficiency of essential nutrients, and fire). All of them have been shaping differently the Neotropical savannah physiognomies, and recent changes in abiotic conditions are occurring in a very rapid manner, mainly due to anthropogenic actions over the last decades. Reduction of native vegetation areas, expansion of agriculture and extensive pecuary, changes in the fire regime and the introduction of nutrients, such as nitrogen and phosphorus, are examples that have been reported for Neotropical savannah in recent times and that may potentially have negative effects on biodiversity. In view of this reality, conservation strategies for the remaining area should be discussed with society and prioritized in government policies in order to conserve such areas and to encourage the sustainable use of the Savannah wealth.

References

  1. 1. Reddy, AR, Chaitanya, KV, Vivekanandan, M. Drought-induced responses of photosynthesis and antioxidant metabolism in higher plants. Journal of Plant Physiology. 2004; 161: 189-1202.
  2. 2. Mittler, R. Abiotic stress, the field environment and stress combination. Trends in Plant Science. 2006; 11: 15-19.
  3. 3. Zavalloni, C, Gielen, B, De Boeck, HJ, Lemmens, CMHM, Ceulemans, R, Nijs, I. Greater impact of extreme drought on photosynthesis of grasslands exposed to a warmer climate in spite of acclimation. Physiologia Plantarum. 2009; 136: 57-72.
  4. 4. Zlatev, Z, Cebola, LF. An overview on drought induced changes in plant growth, water relations and photosynthesis. Emirates Journal of Food and Agriculture. 2012; 24: 57-72.
  5. 5. Nishiyama, Y, Murata, N. Revised scheme for the mechanism of photoinhibition and its application to enhance the abiotic stress tolerance of the photosynthetic machinery. Applied Microbiology Biotechnology. 2014; 98: 8777-8796.
  6. 6. IPCC Climate change 2014. Impacts adaptation and vulnerability. Pat A: Global and sectorial aspects. Contribution of working group II to the fifth assessement report of the intergovernmental panel on climate change. Cambridge: UK. Cambridge University Press. [2014].
  7. 7. Pecl, GT, Araújo, MB, Bell, JD, Blanchard, J, Bonebrake, TC. et al. Biodiversity redistribution under climate change: Impacts on ecosystems and human well-being. Science. 2017; 355: 1-9.
  8. 8. Scarano, FR. Biodiversity Sector: Risks of Temperature Increase to Biodiversity and Ecosystems. In: Nobre C.A.; Marengo J.A.; Soares W.R. [Org.]. Climate Change Risks in Brazil. 1ed.Cham, Switzerland: Springer International Publishing. 2019; pp. 131-141.
  9. 9. Nishiyama R, Watanabe, Y, Leyva-Gonzalez, MA, Van, HC, Fujita, Y, Tanaka, M. et al. Arabidopsis AHP2, AHP3, and AHP5 histidine phosphotransfer proteins function as redundant negative regulators of drought stress response. Proc. Natl. Acad. Sci. U.S.A. 2013; 110: 4840-4845.
  10. 10. Ha, CV, Leyva-Gonzalez, MA, Osakabe, Y, Tran, UT, Nishiyama, R, Watanabe, Y. et al. Positive regulatory role of strigolactone in plant responses to drought and salt stress. Proc. Natl. Acad. Sci. U.S.A. 2014; 111: 581-856.
  11. 11. Myers, N, Mittermeier, RA, Mittermeier, CG, Fonseca, GAB, Ken, J. Biodiversity hotspots for conservation priorities. Nature. 2000; 403: 853-858.
  12. 12. Gottsberger, G, Silberbauer-Gottsberger, I. Life in the cerrado: a South American tropical seasonal ecosystem. Vol. I Origin, Structure, Dynamics and Plant use. Ulm: Reta Verlag. 2006. https://doi.org/10.1086/665973
  13. 13. Mendonça, RC, Felfili, JM, Walter, BMT, Silva-Júnior, MC, Rezende, AB. et al. Flora Vascular do Bioma Cerrado: checklist com 12.356 espécies. In Cerrado: Ecologia e Flora. Vol. 2. Edited by S.M. Sano, S.P. Almeida, and J.F. Ribeiro. Brasília: Embrapa Cerrados. pp. 213-228.
  14. 14. Coutinho, LM. O conceito de Cerrado. Revista Brasileira de Boânica. 1978; 1: 17-2.
  15. 15. Capuzzo JP, Rossatto, DR, Franco, AC. Differences in morphological and physiological leaf characteristics between Tabebuia aurea and T. impetiginosa is related to their typical habitats of occurrence. Acta Botanica Brailica. 2012; 26: 519-526.
  16. 16. Silveira, FAO, Negreiros, D, Barbosa, NPU, Buisson, E, Carmo, FF, Carstensen, DW. et al. Ecology and evolution of plant diversity in the endangered campo rupestre: a neglected conservation priority. Plant Soil. 2016; 403: 29-152.
  17. 17. Franco, AC. Seasonal Patterns of Gas Exchange, Water Relations and Growth of Roupala Montana, an Evergreen Savanna Species. Plant Ecology. 1998; 136: 69-76.
  18. 18. Instituto Nacional de Meteorologia. Ministério da Agricultura, Pecuária e Abastecimento. [Internet]. 2020. Available from: https://portal.inmet.gov.br/
  19. 19. Furley, PA, Ratter, JA. Soil resources and plant communities of the central Brazilian Neotropical savannah and their development. Journal of Biogeography. 1988; 15: 97-108.
  20. 20. Vieira, EA, Silva, MG, Moro, CF, Laura, VA. Physiological and biochemical changes attenuate the effects of drought on the Neotropical savannah species Vatairea macrocarpa (Benth.) Ducke. Plant Physiology and Biochemistry. 2017; 115: 472-483.
  21. 21. Yordanov, I, Velikova, V, Tsonev, T. Plant responses to drought and stress tolerance. Bulg. Journal of Plant Physiology. 2003; Special issue: 187-206.
  22. 22. Chavarria G, Santos HP. Plant Water Relations: Absorption, Transport and Control Mechanisms. In: Montanaro G, Dichio B (eds.) Advances in Selected Plant Physiology Aspects. Rijeka: InTech; 2012. pp. 105-132.
  23. 23. Silva, EC, Albuquerque, MB, Neto, ADA, Júnior, CDS. Drought and Its Consequences to Plants - From Individual to Ecosystem. 2013; Intechopen. http://dx.doi.org/10.5772/53833
  24. 24. Mullet, JE, Whitsitt, MS. Plant cellular responses to water deficit. Plant Growth Regulation. 1999; 20: 119-124.
  25. 25. Blum, A. Drought resistance, water-use efficiency, and yield potential – are they compatible, dissonant, or mutually exclusive? Australian Journal of Agricultural Research. 2005; 56: 1159-1168.
  26. 26. Moore, JP, Vicré-Gibouin, M, Farrant, JM, Driouich, A. Adaptations of higher plant cell walls to water loss: Drought vs. desiccation. Physiologia Plantarum. 2008; 134: 237-245.
  27. 27. Dinakar, C, Djilianov, D, Bartels, D. Photosynthesis in desiccation tolerant plants: Energy metabolism and antioxidative stress defense. Plant Science. 2012; 182: 29-41.
  28. 28. Farrant, JM, Lehner, A, Cooper, K, Wiswedel, S. Desiccation tolerance in the vegetative tissues of the fern Mohria caffrorum is seasonally regulated. Plant Journal. 2009; 57: 65-79.
  29. 29. Franco, AC. Ecophysiology of woody plants. In: Oliveira PS, Marquis RJ [eds] The cerrados of Brazil. Columbia University Press, New York. 2002; pp. 178-197.
  30. 30. Goldstein, G, Meinzer, F, Bucci, SJ, Scholz, G, Franco, AC, Hoffmann, WA. Water economy of Neotropical savanna trees: six paradigms revisited. Tree Physiology. 2008; 28: 395-404.
  31. 31. Rossatto, DR, da Silveira Lobo, SL, Franco, AC. The partitioning of water uptake between growth forms in a Neotropical savanna: do herbs exploit a third water source niche? Plant Biology. 2013; 15: 84-92.
  32. 32. Araya, Y, Silvertown, J, Gowing, DJ, McConway, KJ, Linder, HP, Midgley, G. A fundamental, eco-hydrological basis for niche segregation in plant communities. New Phytologist. 2011; 189: 253-258.
  33. 33. Silvertown, J, Araya, Y, Gowing, D. Hydrological niches in terrestrial plant communities: a review. Journal of Ecology. 2015; 103: 93-108.
  34. 34. Aparecido, LMT, Teodoro, GS, Mosquera, G, Brum, M, Barros, FV, Pompeu, PV. et al. 2017. Ecohydrological drivers of Neotropical vegetation in montane ecosystems. Ecohydrology. 2018;e1932. https://doi.org/10.1002/eco.1932
  35. 35. Castro, SAB, Silveira, FAO, Marcato, MS, Lemos-Filho, JP. So close, yet so different: Divergences in resource use may help stabilize coexistence of phylogenetically-related species in a megadiverse grassland. Flora. 2018; 238: 72-78.
  36. 36. Franco AC, Bustamante M, Caldas LS. et al Leaf functional traits of Neotropical savannah trees in relation to seasonal water deficit. Trees. 2005; 19:326-335.
  37. 37. Bucci, SJ, Goldstei, G, Meinzer, FC, Franco, AC, Campanello, P, Schotz, FG. Mechanisms contributing to seasonal homeostasis of minimum leaf water potential and predawn disequilibrium between soil and plant water potential in Neotropica savanna trees. Trees. 2005; 19: 296-304.
  38. 38. Costa, AC, Rezende-Silva, SL, Megguer, CA, Moura, LMF, Rosa, M, Silva, AA. The effect of irradiance and water restriction on photosynthesis in young jatobá-do-Neotropical savannah (Hymenaea stigonocarpa) plants. Photosynthetica. 2015; 53: 118-127.
  39. 39. França, MGC, Prados, LMZ, Lemos-Filho, JP, Ranieri, BD, Vale, FHA. Morphophysiological differences in leaves of Lavoisiera campos-portoana (Melastomataceae) enhance higher drought tolerance in water shortage events. Journal of Plant Research. 2012; 125: 85-92.
  40. 40. Wolf, S, Greiner, S. Growth control by cell wall pectins. Protoplasma. 2012; 249: 169-175.
  41. 41. Peaucelle, A, Braybrook, S, Höfte, H. Cell wall mechanics and growth control in plants: The role of pectins revisited. Frontier in Plant Science. 2012; 3: 1-6.
  42. 42. Le Gall, H, Phillipe, F, Domon, JM, Gillet, F, Pelloux, J, Rayon, C. Cell Wall Metabolism in Response to Abiotic Stress. Plants. 2015; 4: 112-166.
  43. 43. Filho, JTC, Stacciarini-Seraphin, E. Alteração no potencial osmótico e teor de carboidratos solúveis em plantas jovens de lobeira (Solanum lycocarpum St.-Hil.) em resposta ao estresse hídrico. Revista Brasileira de Botânica. 2001; 24: 199-204.
  44. 44. Tamura, T, Hara, K, Yamaguchi, Y, Koizumi, N, Sano, H. Osmotic stress tolerance of transgenic tobacco expressing a gene encoding a membrane-located receptor-like protein from tobacco plants. Plant Physiology. 2003; 131: 454-62.
  45. 45. Chen, TH, Murata, N. Glycinebetaine: an effective protectant against abiotic stress in plants. Trends in Plant Science. 2008; 13: 499-505.
  46. 46. Silva EC, Nogueira RJMC, Vale FHA, Melo NF, Araujo FP. Water relations and organic solutes production in four umbu tree (Spondias tuberosa) genotypes under intermittent drought. Brazilian Journal of Plant Physiology. 2009; 21(1): 43-53.
  47. 47. Suguiyama, VF, Silva, EA, Meireles, ST, Centeno, DC, Braga, MR. Leaf metabolite profile of the Brazilian resurrection plant Barbacenia purpurea Hook. (Velloziaceae) shows two time dependent response during desiccation and recovering. Frontiers in Plant Science. 2014; 5:1-13.
  48. 48. Souza, LC, Luz, LM, Martins, JTS, Neto, CFO, Palheta, JG, et al. Osmoregulators in Hymenaea courbaril and Hymenaea stigonocarpa under water stress and rehydration. Journal of Forestry Research. 2017; https://doi.org/10.1007/s11676-017-0456-x
  49. 49. Farrant, J, Brandt, WF, Lidsey, GG. An overview of the mechanisms of dessication tolerance in selected angiosperm resurrection plants. Plant Stress. 2007; 1:72-84.
  50. 50. Suguiyama, VF, Sanches, RFE, Meirelles, ST, Centeno, DC, Silva, EA, Braga, MR. Physiological responses to water deficit and changes in leaf cell wall composition as modulated by seasonality in the Brazilian resurrection plant Barbacenia purpurea. South African Journal of Botany. 2016; 105: 270-278.
  51. 51. Mittler, R, Vanderauwera, S, Suzuki, N, Miller, G, Tognetti, VB. et al. ROS signaling: the new wave? Trends in Plant Science. 2011; 16: 300-309.
  52. 52. Del Río, LA. ROS and RNS in plant physiology: an overview. Journal of Experimental Botany. 2015; 66: 2827-2837.
  53. 53. Turkey, I. ROS and RNS: key signalling molecules in plants. Journal of Experimental Botany. 201; 69:3313-3315.
  54. 54. Dietz, KJ, Turkan, I, Krieger-Liszkay, A. Redox- and reactive oxygen species-dependent signaling in and from the photosynthesizing chloroplast. Plant Physiology. 2016; 171:1541-1550.
  55. 55. Silva, EC, Albuquerque, MB, Neto, ADA, Junior, CDS. Drought and Its Consequences to Plants – From Individual to Ecosystem. In: Responses of Organisms to Water Stress. Intech. pp. 17-47. 2013.
  56. 56. Turkan I. Emerging roles for ROS and RNS—versatile molecules in plants. Journal of Experimental Botany. 2017; 68: 4413-4416.
  57. 57. Assaha, DV, Liu, L, Ueda, A, Nagaoka, T, Saneoka, HEffects of drought stress on growth, solute accumulation and membrane stability of leafy vegetable, huckleberry (Solanum scabrum Mill.). Journal of Environmental Biology. 2016; 37: 107e114.
  58. 58. Vieira, EA, Galvão, FCA, Barros, AL. Influence of water limitation on the competitive interaction between two Neotropical savannah species and the invasive grass Brachiaria brizantha cv. Piata. Plant Physiology and Biochemistry. 2019; 135: 206-214.
  59. 59. Mollo, L, Martins, MCM, Oliveira, VF, Nievola, CC, Figueiredo-Ribeiro, RCL. Effects of low temperature on growth and non-structural carbohydrates of the imperial bromeliad Alcantarea imperialis cultured in vitro. Plant Cell, Tissue Organ Culture. 2011; 107: 141-149.
  60. 60. Mathur S, Agrawal D, Jajoo A. Photosynthesis: Response to high temperature stress. Journal Photochemistry Photobiology B Biol. 2014; 137: 116-126.
  61. 61. Petrov, V, Hille, J, Mueller-Roeber, B, Gechev, TS. ROS-mediated abiotic stress-induced programmed cell death in plants. Frontier Plant Science. 2015; 69: 1-16.
  62. 62. Lemos-Filho, JP. Photoinhibition in three species of the savanna (Annona crassifolia, Eugenia dysenterica and Campomanesia adamantium) in the dry and rainy season. Revista Brasileira de Botânica. 2000; 23: 45-50.
  63. 63. Lüttge, U. Duarte, HM, Scarano, FR, Mattos, EA, Cavalin, PO. et al. Physiological ecology of photosynthesis of five sympatric species of Velloziaceae in the rupestrian fields of Serra do Cipó, Minas Gerais, Brazil. Flora. 2007; 202: 637-646.
  64. 64. Morales, M, Garcia, QS, Silva, AIS, Silva, MC, Bosch, SM. Tocotrienols in Vellozia gigantea leaves: occurrence and modulation by seasonal and plant size effects. Planta. 2014; 240: 437-446.
  65. 65. Castro, SAB, Sá, CEM, Mourão, FA, Duarte, HM, Fernandes, GW, Lemos-Filho, JP. Ecophysiological performance of a threatened shrub under restored and natural conditions in harsh tropical mountain environments. Acta Botânica Brasílica. 2016; 30: 17-26.
  66. 66. Grassi, G, Magnani, F. Stomatal, mesophyll conductance and biochemical limitations to photosynthesis as affected by drought and leaf ontogeny in ash and oak trees. Plant, Cell & Environment. 2005; 28: 834-849.
  67. 67. Kuster, VC, Castro, SAB, Vale, FHA. Environmental conditions modulate plasticity in the physiological responses of three plant species of the Neotropical savannah. Acta Physiology Plantarum. 2017; 39: 103.
  68. 68. Franco, AC, Lüttge, U. Midday depression in savanna trees: coordinated adjustments in photochemical efficiency, photorespiration, CO2 assimilation and water use efficiency. Oecologia. 2002; 131:3 56-365.
  69. 69. Morales, M, Garcia, QS, Bosch, SM. Ecophysiological response to season variations in water availability in the arborescent, endemic plant Vellozia gigantea. Tree Physiology. 2015; 35: 253-265.
  70. 70. Ducruet, JM, Peeva, V, Havaux, M. Chlorophyll thermofluorescence and thermoluminescence as complementary tools for the study of temperature stress in plants. Photosynthesys Research. 2007; 93: 159-171.
  71. 71. Gian-Reto, W, Post, E, Convey, P, Menzel, A, Parmesan, C, BeeBee, TJC. et al. Ecological responses to recent climate change. Nature. 2002; 416: 389-395.
  72. 72. Corlett RT, Westcott DA. Will plant movements keep up with climate change? Trends in Ecology & Evolution. 2013; 28(8): 482-488.
  73. 73. Melo, NMJ, Rosa, RSEG, Pereira, EG, Souza, JP. Rising (CO2) changes competition relationships between native woody and alien herbaceous Cerrado species. Functional Plant Biology. 2018; 45(8): 854-864.
  74. 74. Souza, JP, Melo, NMJ, Halfeld, AD, Vieira, KIC, Rosa, BL. Elevated atmospheric CO2 concentration improves water use efficiency and growth of a widespread Cerrado tree species even under soil water déficit. Acta Botanica Brasilica. 2019; 33(3): 425-436.
  75. 75. Yamori, W, Hikosaka K, Way, DA. Temperature response of photosynthesis in C3, C4, and CAM plants: temperature acclimation and temperature adaptation. Photosynthesis Research. 2014; 119:101-117.
  76. 76. Wahid A, Gelani S, Ashraf M, Foolad MR. Heat tolerance in plants: An overview. Environmental Experimental Botany. 2007; 61 (3): 199-223.
  77. 77. Slot, M, Winter, K. In situ temperature response of photosyn- thesis of 42 tree and liana species in the canopy of two Panamanian lowland tropical forests with contrasting rainfall regimes. New Phytologist. 2017; 214: 1103-1117.
  78. 78. Menezes-Silva, PE, Lourenço, LL, Alves, RDFB, Souza, LF, Almeida, SES, Farnese, FS. Different ways to die in a changing world: Consequences of climate change for tree species performance and survival through an ecophysiological perspective. Ecology and Evolution. 2019; https://doi.org/10.1002/ece3.5663
  79. 79. Franco, AC, Rossatto, DR, Silva, LCR, Ferreira, CS. Cerrado vegetation and global change: the role of functional types, resource availability and disturbance in regulating plant community responses to rising CO2 levels and climate warming. Theoretical Experimental Plant Physiology. 2014; 26: 19-38.
  80. 80. Salvucci, ME, Crafts-Brandner, SJ. Relationship between the heat tolerance of photosynthesis and the thermal stability of rubisco activase in plants from contrasting thermal environments. Plant Physiology. 2004; 134: 1460-1470.
  81. 81. Scafaro, A.P, Gallé, A, Rie AV, Carmo-Silva, E, Salvucci, ME, Atwell, BJ. Heat tolerance in a wild Oryza species is at- tributed to maintenance of Rubisco activation by a thermally stable Rubisco activase ortholog. New Phytologist. 2016; 211: 899-911.
  82. 82. Zhang, M, Chen, Q , Shen, S. Physiological responses of two Jerusalem artichoke cultivars to drought stress induced by polyethylene glycol. Acta Physiologia Plantarum. 2011; 33: 313-318.
  83. 83. Cui, M, Lin, Y, Zu, Y, Efferth, T, Li, D, Tang, Z. Ethylene Increases Accumulation of Compatible Solutes and Decreases Oxidative Stress to Improve Plant Tolerance to Water Stress in Arabidopsis. Journal Plant Biology. 2015; 58: 193-201.
  84. 84. Nievola, CC, Carvalho, CP, Carvalho, V, Rodrigues, E. Rapid responses of plants to temperature changes. Temperature. 2017; 4: 371-405.
  85. 85. Schrader, SM, Wise, RR, Wacholtz, WF, Ort, DR. Thylakoid membrane responses to moderately high leaf temperature in Pima cotton. Plant Cell Environmental. 2004; 27: 725-735.
  86. 86. Tang T, Liu P, Zheng G, Li W. Two phases of response to long-term moderate heat: Variation in thermotolerance between Arabidopsis thaliana and its relative Arabis paniculata. Phytochemistry. 2016;122: 81-90.
  87. 87. Reda, F, Mandoura, HMH. Response of enzymes activities, photosynthetic pigments, proline to low or high temperature stressed wheat plant (Triticum aestivum L.) in the presence or absence of exogenous proline or cysteine. International Journal Academic Research. 2011; 3: 108-115.
  88. 88. Vierling, E. The roles of heat shock proteins in plants. Annual Review Plant Physiology and Plant Molecular Biology. 1991; 42: 579-620.
  89. 89. Hemantaranjan, A, Bhanu, AN, Singh, MN, Yadav, DK. et al. Heat Stress Responses and Thermotolerance. Advances in Plants & Agriculture Research. 2014; 1(3): 00012.
  90. 90. Ohama N, Sato H, Shinozaki K, Yamaguchi-Shinozaki K. Transcriptional regulatory network of plant heat stress response. Trends Plant Sci. 2017; 22(1): 53-65.
  91. 91. Duarte, AA, Silva, CJ, Marques, AR, Modolo, Lv, Lemos-Filho, JP. Does oxidative stress determine the thermal limits of the regeneration niche of Vriesea friburgensis and Alcantarea imperialis (Bromeliaceae) seedlings? Journal of Thermal Biology. 2019; https://doi.org/10.1016/j.jtherbio.2019.02.003
  92. 92. Franco, AC, Matsubara, S, Orthen, B. Photoinhibi-tion, carotenoid composition and the co-regulation of photochemical and non-photochemical quenching in Neotropical savannah trees. Tree Physiology. 2008; 27: 717-725
  93. 93. Nabout, JC. Global warming decreases the morphological traits of germination and environmental suitability of Dipteryx alata (Fabaceae) in Brazilian Cerrado. Acta Botânica Brasilica. 2019; 33: 446-453.
  94. 94. Gomes, MP, Brito, JCM, Silva, JG, Cruz, FVS, Bicalho, EM. et al. Temperature effects on Zn-responses and Zn-reclamation capacity of two native Brazilian plant species: Implications of climate change. Environmental and Experimental Botany. 2018; 155: 589-599.
  95. 95. Drake, JE, Vårhammar, A, Kumarathunge, D, Medlyn, BE, Pfautsch, S, Reich, PB. et al. A common thermal niche among geographically diverse populations of the widely distributed tree species Eucalyptus tereticornis: No evidence for adaptation to climate- of- origin. Global Change Biology. 2017; 23: 5069-5082.
  96. 96. Faria, AP, Marabesi, MA, Gaspar, M, França, MGC. The increase of current atmospheric CO2 and temperature can benefit leaf gas exchanges, carbohydrate content and growth in C4 grass invaders of the Cerrado biome. Plant Physiology and Biochemistry. 2018; 127: 608-616.
  97. 97. Leadley, P, Proença, V, Fernández-Manjarrés, Pereira, HM, Alkemade, R. et al. Interacting regional-scale regime shifts for biodiversity and ecosystem services. Bioscience. 2014; 64: 893-905.
  98. 98. Pamesan, C, Hanley, ME. Plants and climate change: complexities and surprises. Annals of Botany. 2015; 116: 849-864.
  99. 99. Oliveira, U, Soares-Filho, BS, Paglia, AP, Brescovit, AD, Carvalho, CJB. et al. Biodiversity conservation gaps in the Brazilian protected areas. Scientific Reports. 2017; 7: 1-9.
  100. 100. Reis, CR, Bueno, ML, Rocha, LF, Santos, LG, Gorgens, EB. Climate Change Influencing the Potential Distribution of a Brazilian Savanna Indicator Species. Floresta e Ambiente. 2019; 26[3]: e20180085.
  101. 101. Barbosa, ILBD, Oliveira, HR, Terribile, LC, Diniz-Filho, JAF. Geographical distribution of Stryphnodendron adstringens Mart. Coville [Fabaceae]: modeling effects of climate change on past, present and future. Brazilian Jornal of Botany. 2019; 42: 53-61.
  102. 102. Chaves, CJ, Leal, BSS, Lemos-Filho, JP. How are endemic and widely distributed bromeliads responding to warming temperatures? A case study in a Brazilian hotspot. Flora. 2018; 238: 110-118.
  103. 103. Silva, JMC, Rapini, A, Barbosa, LCF, Torres, RR. Extinction risk of narrowly distributed species of seed plants in Brazil due to habitat loss and climate change. Peer Journal. 2019; https://doi.org/10.7717/peerj.7333
  104. 104. Bueno, ML, Pennington, RT, Dexter, KG, Kamiro, LHY, Pontara, V. et al. Effects of Quaternary Climatic Fluctuations on the Distribution of Neotropical Savanna Tree Species. Ecography. 2017; 40: 403-414.
  105. 105. Keppel, G, Niel, KPV, Wardell-Johnson, GW, Yates, CJ, Byrne, M. et al. Refugia: identifying and understanding safe havens for biodiversity under climate change. Global Ecology Biogeography. 2012; 21: 393-404.
  106. 106. Collevatti, RG, Ribeiro, MSL, Souza Neto, AC, Franco, AA, Oliveira, G, Terribile, LC. Recovering the demographical history of a Brazilian Cerrado tree species Caryocar brasiliense: coupling ecological niche modeling and coalescent analyses. Natureza e Conservação. 2012; 10: 169-176.
  107. 107. Gómez, JM, Valladares, F, Puerta-Pinero, C. Differences between structural and functional environmental heterogeneity caused by seed dispersal. Functional Ecology. 200; 18: 787-792.
  108. 108. Chen, FS, Zeng, DH, Fahey, TJ, Yao, CY, Yu, ZY. Response of leaf anatomy of Chenopodium acuminatum to soil resource availability in a semi-arid grassland. Plant Ecology. 2010; 209: 375-382.
  109. 109. Valladares, F, Gianoli, E, Gómez JM Ecological limits to plant phenotypic plasticity. New Phytologyst. 2007; 176: 749-763.
  110. 110. Poorter, H, Niinemets, U, Ntagkas, N, Siebenkas, A, Mäenpää et al. A meta-analysis of plant responses to light intensity for 70 traits ranging from molecules to whole plant performance. New Phytologist. 2019; 223: 1073-1105.
  111. 111. Barderrama, SIV, Chazdon, RL. Light-dependent seedling survival and growth of four tree species in Costa Rican second-growth rain forests. Journal of Tropical Ecology. 2005; 21: 383-395.
  112. 112. Rodrıguez-Calcerrada, J, Cano, FJ, Valbuen-Carabanã, M, Aranda, LG. Functional performance of oak seedlings naturally regenerated across microhabitats of distinct overstorey canopy closure. New Forest. 2010; 39: 245-259.
  113. 113. Valladares, F, Arrieta, S, Aranda, I, Lorenzao, D, Sanchez, D. et al. Shade tolerance, photoinhibition sensitivity and phenotypic plasticity of Ilex aquifolium in continental Mediterranean sites. Tree Physiol. 2005; 25: 1041-1052.
  114. 114. Demmig-Adams B, Adams, WW. The role of xanthophyll cycle carotenoids in the protection of photosynthesis. Trends in Plant Science. 1996; 1: 21-26.
  115. 115. Kouril, R, Wientjes, E, Bultema, JB, Croce, R, Boekema, EJ. High-light vs. lowlight: effect of light acclimation on photosystem II composition and organization in Arabidopsis thaliana. Biochim Biophys Acta. 2013; 1827: 411-419.
  116. 116. Cookson, SJ, Granier, C. A dynamic analysis of the shad induced plasticity in Arabidopsis thaliana rosette leaf development reveals new components of the shade-adaptative response. Annals of Botany. 2006; 97: 443-452.
  117. 117. Valladares F, Niinemets, Ü. Shade tolerance, a key plant feature of complex nature and consequences. Annual Rewiew of Ecology, Evolution and Systematics. 2008; 39:237-257.
  118. 118. Rooijen, R, Aarts, MGM, Harbinson, J. Natural genetic variation for acclimation of photosynthetic light use efficiency to growth irradiance in Arabidopsis. Plant Physiology. 2015; 167:1412-1429.
  119. 119. Ronquim, CS, Prado, CHBA, Souza, JP. Growth, photosynthesis and leaf water potential in young plants of Copaifera langsdorffii Desf. [Caesalpiniaceae] under contrasting irradiances. Brazilian Society of Plant Physiology. 2009; 21: 197-208.
  120. 120. Barros, FV, Goulart, MF, Telles, SBS, Lovato, MB, Valladares, F, Lemos-Filho, JP. Phenotypic plasticity to light of two congeneric trees from contrasting habitats: Brazilian Atlantic Forest versus cerrado [savanna]. Plant Biology. 2010. doi:10.1111/j.1438-8677.2011.00474.x
  121. 121. Rossatto, DV, Araújo, PE, Silva, BHP, Franco AC. Photosynthetic responses of understory savanna plants: Implications for plant persistence in savannas under tree encroachment. Flora. 2018; https://doi.org/10.1016/j.flora.2017.12.009
  122. 122. Goulart, MF, Lovato, MB, Barros, FV, Valladares, F, Lemos-Filho, JP. Which Extent is Plasticity to Light Involved in the Ecotypic Differentiation of a Tree Species from Savanna and Forest? Biotrópica. 2011; DOI 10.1111/j.1744-7429.2011.00760.x
  123. 123. Costa, AC, Rezende-Silva, SL, Megguer, CA, Moura, LMF, Rosa, M, Silva, AA. The effect of irradiance and water restriction on photosynthesis in young jatobá-do-cerrado (Hymenaea stigonocarpa) plants. Photosynthetica. 2015; 53(1): 118-127.
  124. 124. Pinheiro, LFS, Kolb, RM, Rossatto, DR. Changes in irradiance and soil properties explain why typical non-arboreal savanna species disappear under tree encroachment. Australian Journal of Botany. 2016; 64: 333-341.
  125. 125. Lepsch, IF. Cerrado: Soils from the. Encyclopedia of Soil Science, Thirt Edition. Taylor & Francis. 2016; p. 1-3. DOI: 10.1081/E-ESS3-120053880
  126. 126. Lopes, AJ, Cox, FR. A survey of the fertility status of surface soils under cerrado vegetation of Brazil. Soil Science Society of America Journal. 1977; 41: 752-757.
  127. 127. Haridasan, M. Nutritional adaptations of native plants of the cerrado biome in acid soils. Brazilian Journal of Plant Physiology. 2008; 20(3): 183-195.
  128. 128. Ribeiro, JF, Walter, BMT. Fitofisionomias do bioma Cerrado. In S. M. Sano, & S. P. Almeida (Eds.) Cerrado: Ecologia e flora Planaltina, DF: EMBRAPA-CPAC. p. 151-212. 2008.
  129. 129. FAO. Soil Map of the World: 1: 500.000, Legend. UNESCO: Paris, 1974; 1 pp.
  130. 130. Reatto, A, Correia, JR, Spera, ST. Solos do bioma cerrado: aspectos pedológicos. In: Sano, SM; Almeida, S.P. de, ed. Cerrado: ambiente e flora. Planaltina: EMBRAPA-CPAC, 1998. p. 47-86.
  131. 131. Lambers, H, Costa, PB, Oliveira, RS, Silveira, FAO. Towards more sustainable cropping systems: lessons from native Cerrado species Hans Lambers Theoretical Experimental of Plant Physiology. 2020; https://doi.org/10.1007/s40626-020-00180-z
  132. 132. Araújo, GM, Haridasan, M. A comparison of the nutritional status of two forests on dystrophic and mesotrophic soils in the cerrado region of central Brazil. Communications in Soil Science and Plant Analysis. 1988; 19: 1075-1089.
  133. 133. Neri, AV, Schaefer, CEGR, Silva, AF, Souza, AL, Ferreira-Junior, WG, Meira-Neto, JAA. The influence of soils on the floristic composition and community structure of an area of brazilian cerrado vegetation. Edinburgh Journal of Botany. 2012; 69(1): 1-27.
  134. 134. Delhaize, E, Ryan, PR. Aluminum toxicity and tolerance in plants. Plant Physiology. 1995; 107: 315-321.
  135. 135. Poschenrieder, C, Gunsé, B, Corrales, I, Barceló, J. A glance into aluminum toxicity and resistance in plants. Science Total Environmental. 2008; 400: 356-368.
  136. 136. Brunner, I, Sperisen, C. Aluminum exclusion and aluminum tolerance in woody plants. Frontier Plant Scirence. 2013; 1: 172.
  137. 137. von Uexkull, HR, Mutert, E. Global extent, development and economic impact of acid soils. In: Date RA, et al. (eds) Plant soil interactions at low pH. Kluwer Academic Publishers, Dordrecht. 1995; pp 5-19.
  138. 138. Kinraide, T.B. Reconsidering the rhizotoxicity of hydroxyl, sulphate, and fluoride complexes of Aluminum. Journal of Experimental Botany. 1997, 48: 1115-1124.
  139. 139. Siqueira-Neto, Piccolo, MC, Scopel, E, Junior, CC, Cerri, CC, Bernoux, M. Carbono total e atributos químicos com diferentes usos do solo no cerrado. Acta Scientiarum Agronomy. 2009; 31: 709-717.
  140. 140. Bojórquez-Quintal, E, Escalante-Magaña, C, Echevarría-Machado, I, Martínez-Estévez, M. Aluminum, a friend or foe of higher plants in acid soils. Frontier Plant Science. 2017; 8:1-18.
  141. 141. Singh, SK, Reddy, VR, Fleisher, DH, Timlin, DJ. Relationship between photosynthetic pigments and chlorophyll fluorescence in soybean under varying phosphorus nutrition at ambient and elevated CO2. Photosynthetica. 2017; 55: 421-433.
  142. 142. Souza, MC, Habermann, G, do Amaral, CL, Rosa, AL, Pinheiro, MHO, da Costa, FBVochysia tucanorum Mart.: an aluminum-accumulating species evidencing calcifuge behavior. Plant Soil. 2017; 419: 377-389.
  143. 143. Haridasan, M. Aluminum accumulation by some cerrado native species of central Brazil. Plant Soil. 1982; 65: 265-273.
  144. 144. Jansen, S, Broadley, MR, Robbrecht, E, Smets, E. Aluminum hyperaccumulation in angiosperms: a review of its phylogenetic significance. The Botanical Review. 2002; 68: 235-269.
  145. 145. Souza, MC, Bueno, PCP, Morellato, LPC, Habermann, G. Ecological strategies of Al-accumulating and non-accumulating functional groups from cerrado sensu stricto. Anais Academia Brasileira Ciências. 2015; 87(2): 813-823.
  146. 146. Kochian, LV, Piñeros, MA, Liu, J, Magalhaes, JV. Plant adaptation to acid soils: the molecular basis for crop Aluminum resistance. Annual Review Plant Biology. 2015; 66: 571-598.
  147. 147. Alvin, MN, França, MGC, Ramos, FT, Nascentes, CC. Slight soil deacidification compromisse the growth and the aluminum accumulation in Qualea cordata plants. Applied Ecology and Environmental Research. 2017; 15(4):2023-2034.
  148. 148. Cury, NF, Silva, RCC, Andre, MSF, Fontes, W, Ricart, CAO. et al. Root proteome and metabolome reveal a high nutritional dependency of Aluminum in Qualea grandiflora Mart. (Vochysiaceae). Plant Soil. 2020; 446:125-143 https://doi.org/10.1007/s11104-019-04323-3
  149. 149. Andrade, LRM, Barros, LMG, Echevarria, GF, Amaral, LIV, Cotta, MG. et al. Al-hyperaccumulator Vochysiaceae from the Brazilian Cerrado store aluminum in their chloroplasts without apparent damage. Environmental and Experimental Botany. 2011; 70: 37-42.
  150. 150. Banhos, OFAA, Souza, MC, Habermann, G. High aluminum availability may affect Styrax camporum, an Al non-accumulating species from the Brazilian savanna. Theoretical and Experimental Plant Physiology. 2016; https://doi 10.1007/s40626-015-0051-5
  151. 151. Bressan, ACG, Silva, GS, Banhos, OFAA, Tanaka, FAO, Haermann, G. Physiological, anatomical and ultrastructural effects of aluminum on Styrax camporum, a native Cerrado woody species. Journal of Plant Research. 2020; https://doi.org/10.1007/s10265-020-01210-2.
  152. 152. Čiamporová, M. Morphological and structural responses of plant roots to Aluminum at organ, tissue, and cellular levels. Biologia Plantarum. 2002; 45: 161-171.
  153. 153. Silva, CO, Brito, DS, Silva, AA, Rosa, VR, Santos, MFS. et al. Diferential accumulation of aluminum in root tips of soybean seedlings. Brazilian Journal of Botany. 2020; 43: 99-107.
  154. 154. Matsumoto, H. Cell Biology of Aluminum Toxicity and Tolerance in Higher Plants. International review of cytology. 2000; 200: 1-46.
  155. 155. Bittencourt, BMOC, Silva, CMS, Filho, SZ, Habermann, G. Aluminum (Al)-induced organic acid exudation in an Al-Accumulating species from the Brazilian savanna. Trees. 2019; 34: 155-162.
  156. 156. Mendonça, AMC, Lira, DA, Vilela, ALO, Vieira, DA, Melo, NC, Barbosa, JPRAD. High aluminum concentration and initial establishment of Handroanthus impetiginosus: clues about an Al non-resistant species in Brazilian Cerrado. Journal of Forestry Research. 2019; https://doi.org/10.1007/s11676-019-01033-5
  157. 157. Konrad, MLF, Silva JAB, Furlani, PR, Machado, EC. Gas exchange and chlorophyll fluorescence in six coffee cultivars under aluminum stress. Bragantia. 2005: 64: 339-347.
  158. 158. Silva, S, Pinto, G, Dias, MC, Correia, CM, Moutinho-Pereira, J, et al. Aluminum long-term stress differently affects photosynthesis in rye genotypes. Plant Physiology Biochemistry. 2012; 54: 105-112.
  159. 159. Vitorello, VA, Capaldi, FR, Stefanuto, VA. Recent advances in aluminum toxicity and resistance in higher plants. Brazilian Journal Plant Physiology. 2005; 17: 129-143.
  160. 160. Chen, LS, Qi, YP, Smith, BR, Liu, XH. Aluminum-induced decrease in CO2 assimilation in citrus seedlings is unaccompanied by decreased activities of key enzymes involved in CO2 assimilation. Tree Physiology. 2005; 25: 317-324.
  161. 161. Chen, LS, Qi, YP, Liu, XH. Effects of aluminum on light energy utilization and photoprotective systems in citrus leaves. Annals of Botany. 2005; 96: 35-41.
  162. 162. Guo, P, Yi-Ping, Q , Yan-Tong, C, Tao-Yu, Y, Lin-Tong, Y. et al. Aluminum effects on photosynthesis, reactive oxygen species and methylglyoxal detoxification in two Citrus species differing in Aluminum tolerance. Tree Physiology. 2018; 38: 1548-1565.
  163. 163. Peixoto, PH, Da Matta, FM, Cambraia, J. Responses of the photosynthetic apparatus to aluminum stress in two sorghum cultivars. Journal Plant Nutrition. 2002; 25: 821-832.
  164. 164. Pereira, WE, Siqueira, DL, Martinez, CA, Puiatti, M. Gas exchange and chlorophyll fluorescence in four citrus rootstocks under aluminum stress. Journal Plant Physiology. 2000; 157: 513-520.
  165. 165. Moustakas, M, Ouzounidou, G, Lannoye, R. Aluminum effects on photosynthesis and elemental uptake in an aluminum-tolerant and non-tolerant wheat cultivar. Journal Plant Nutrition. 1995; 18: 669-683.
  166. 166. Bustamante, MMC, Medina E, Asner GP, Nardoto GB, Garcia-Montiel, DC. et al. Nitrogen cycling in tropical and temperate savannas. Biogeochemistry. 2006; 79: 209-237.
  167. 167. Elser, JJ, Bracken, MES, Cleland, EE, Gruner, D, Stanley, W. et al. Global analysis of nitrogen and phosphorus limitation of primary producers in freshwater, marine and terrestrial ecosystems. Ecology Letters. 2007; 10: 1135-1142.
  168. 168. Lambers, H, Raven, J A, Shaver, GR, Smith, SE. Plant nutrient-acquisition strategies change with soil age. Trends Ecology Evolution. 2008; 23: 95-103.
  169. 169. Kamble, BM, Todmal, SM. Effect of different levels and sources of nitrogen on soil properties, nutrient uptake, and yield of wheat grown on saline soil. International Journal of Chemical Studies. 2020; 8(1): 687-693.
  170. 170. Plaxton, WC, Tran, HT. Metabolic adaptations of phosphate-starved plants. Plant Physiol. 2011; 156: 1006-1015.
  171. 171. Havlin, JL, Tisdale, SL, Nelson, WL, Beaton, JD. Phosphorus. In Soil Fertility and Fertilizers, 8th ed.; Lawrensen, W., Gohn, J., Eds.; Pearson Inc.: Upper Saddle River, NJ, USA, 2014; pp. 185-221.
  172. 172. Pallardy, SG. Nitrogen Metabolism. In Physiology of Woody Plants, 3rd ed.; Pallardy, S.G., Ed.; Academic Press: Missouri, MO, USA, 2008; pp. 233-254.
  173. 173. Reis, DN, Silva, FG, Santana, RC, Oliveira, TC, Freiberger, MB. et al. Growth, Physiology and Nutrient Use Efficiency in Eugenia dysenterica DC under Varying Rates of Nitrogen and Phosphorus. Plants. 2020; 9: 722.
  174. 174. Lopes, AS, Guilherme, LRG. A career perspective on soil management in the Cerrado region of Brazil. Advances in Agronomy. 2016; 137: 1-72.
  175. 175. Viani, RAG, Rodrigues, RR, Dawson, TE, Oliveira, RS. Savanna soil fertility limits growth but not survival of tropical forest tree seedlings. Plant Soil. 2011; 349:341-353.
  176. 176. Pereira, CG, Clode, PL, Oliveira, RS, Lambers, H. Eudicots from severely phosphorus-impoverished environments preferentially allocate phosphorus to their mesophyll. New Phytologist. 2017; 218: 959-973.
  177. 177. Machado, RB, Ramos Neto, MB, Pereira, P, Caldas, EF, Gonçalves DA, et al. Estimativas de Perda da Área do Cerrado Brasileiro. Conservation International do Brasil, Brasília. 2004. pp. 25.
  178. 178. Hunke, P, Mueller, EN, Schröder, B, Zeilhofer, P. The Brazilian Cerrado: assessment of water and soil degradation in catchments under intensive agricultural use. Ecohydrology. 2015; 8: 1154-1180.
  179. 179. Jacobson, TKB, Bustamante, MMC, Kozovits, AR. Diversity of shrub tree layer, leaf litter decomposition and N release in a Brazilian Cerrado under N, P and N plus P additions. Environmental Pollution. 2011; 159: 2236-2242.
  180. 180. Bustamante, MMC, Brito, DQ , Kozovitz, AR, Luedemann, G, Mello, TRB. et al. Effects of nutrient additions on plant biomass and diversity of the herbaceous-subshrub layer of a Brazilian savanna (Cerrado). Plant Ecology. 2012; 213: 795-808.
  181. 181. Lannes, LS, Bustamante, MMC, Edwards, PJ, Venterink, HO. Native and alien herbaceous plants in the Brazilian Cerrado are (co-) limited by different nutrients. Plant Soil. 2015; DOI 10.1007/s11104-015-2725-9
  182. 182. Lambers, H, Brundrett, MC, Raven, JA, Hopper SD. Plant mineral nutrition in ancient landscapes: high plant species diversity on infertile soils is linked to functional diversity for nutritional strategies. Plant Soil. 2010; 334: 11-31.
  183. 183. Laliberté, E, Grace, JB, Huston, MA, Lambers, H, Teste, FP. et al. How does pedogenesis drive plant diversity? Trends in Ecology & Evolution. 2013; 28: 331-340.
  184. 184. Pausas, JG, Keeley, JE. A burning story: the role of fire in the history of life. Bioscience. 2009; 59: 593-601.
  185. 185. Bond, WJ, Midgley, JJ. Fire and the angiosperm revolutions. International Journal Plant Science. 2012; 173: 569-583.
  186. 186. Miranda, HS, Bustamante, MM, Miranda, AC. The fire factor. In: Oliveira, P.S.; Marquis, R.J. The Cerrados of Brazil. New York: Columbia University Press, 2002; pp. 51-68.
  187. 187. Fidelis, A. Is fire always the “bad guy”? Flora. 2020; 268: 151611.
  188. 188. Berlinck, CN, Batista, EKL. Good fire, bad fire: It depends on who burns. Flora. 2020; 268: 1-4.
  189. 189. Moraes, MG, Carvalho, MAA, Franco, AC, Pollock, CJ, Figueiredo-Ribeiro, RCL. Fire and Drought: Soluble Carbohydrate Storage and Survival Mechanisms in Herbaceous Plants from the Cerrado. BioScience. 2016; 66: 107-117.
  190. 190. Bowman, DMJS, Balch. J, Artaxo, P, Bond, WJ, Cochrane, M.A. et al. The human dimension of fire regimes on Earth Journal Biogeography. 2011; 38: 2223-2236.
  191. 191. Hoffmann, W.A. Fire and population dynamics of woody plants in a neotropical savanna: matrix model projections. Ecology. 1999; 80: 1354-1369.
  192. 192. Keeley, JE, Pausas, JG, Rundel, PW, Bond, WJ, Bradstock, RA. Fire as an evolutionary pressure shaping plant traits. Trends in Plant Science. 2011; 16: 406-411.
  193. 193. Fahn, A. and Cutler, D. Xerophytes. Gebruder Borntraeger, Berlin. 1992, Pp. 176.
  194. 194. Loram-Lourenço, L, Farnese, FDS, Sousa, LFD, Alves, RDFB, Andrade, MCP. et al. A Structure Shaped by Fire, but Also Water: Ecological Consequences of the Variability in Bark Properties Across 31 Species From the Brazilian Cerrado. Frontier Plant Science. 2020; 10: 1718.
  195. 195. Dairel, M, Fidelis, A. How does fire affect germination of grasses in the Cerrado? Seed Science Research. 2020; https://doi.org/10.1017/S0960258520000094
  196. 196. Coutinho, LM. Ecological effects of fire in Brazilian Cerrado. Berlin, Heidelberg, Springer. 1982; pp. 273-291.
  197. 197. Soares, JJ, Souza, MHAO, Lima, MIS. Twenty years of post-fire plant succession in a "cerrado", São Carlos, SP, Brazil. Brazilian Journal of Biology. 2006; 66
  198. 198. Felfili, JM, Silva-Junior, MC, Dias, BJ, Rezende, AV. Estudo fenológico de Stryphnodendron adstringens (Mart.) Coville no cerrado sensu stricto da Fazenda Água Limpa no Distrito Federal, Brasil. Revista Brasileira Botânica. 1999; 22(1): 83-90.
  199. 199. Silvério, DV, Pereira, OR, Mews, HA, Maracahipes-Santos, L. Santos, JO, Lenza, E. Surface fire drives short-term changes in the vegetative phenology of woody species in a Brazilian savanna. Biota Neotropical. vol. 15 no. 3 Campinas. 2015. Epub Sep 18.
  200. 200. Pilon, NAL, Hoffmann, WA, Abreu, RCR, Durigan, G. Quantifying the short-term flowering after fire in some plant communities of a cerrado grassland. Plant Ecology & Diversity. 2018; DOI: 10.1080/17550874.2018.1517396
  201. 201. Munhoz, C, Felfili, JM. Reproductive phenology of an herbaceous sub-shrub layer of a Savannah (Campo Sujo) in the Cerrado Biosphere Reserve I, Brazil. Brazilian Journal of Biology. 2007; 67: 299-307.
  202. 202. Neves, DRM, Damasceno-Junior, GA. Post-fire phenology in a campo sujo vegetation in the Urucum plateau, Mato Grosso do Sul, Brazil. Brazilian Journal Biology. 2011; 71: 881-888.
  203. 203. Fidelis, A, Rosalem, P, Zanzarini, V, Camargos, LS, Martins, AR. From ashes to flowers: a savanna sedge initiates flowers 24 h after fire. Ecology. 2019; 100: e02648.
  204. 204. Lamont, BB, Downes, KS. Fire-stimulated flowering among resprouters and geophytes in Australia and South Africa. Plant Ecology. 2011; 212: 2111-2125.
  205. 205. Clarke, PJ, Lawes, MJ, Midgley, JJ, Lamont, BB, Ojeda, F, Burrows, GE, Enright, NJ, Knox, KJEResprouting as a key functional trait: How buds, protection and resources drive persistence after fire. New Phytologist. 2013; 197: 19-35.
  206. 206. Appezzato-da-Glória, B, Cury, G, Soares, MKS. Underground systems of Asteraceae species from the Brazilian Cerrado. Journal of the Torrey Botanical Society. 2008; 135(1): 103-113.
  207. 207. Palhares, D, Paula, JE, Pereira, LAR, Silveira, CES. Comparative wood anatomy of stem, root and xylopodium of Brosimum gaudichaudii (Moraceae) IAWA Journal, 2007; 28(1): 83-94.
  208. 208. Braga, MR, Carpita, NC, Dietrich, SMC, Figueiredo-Ribeiro, RCL. Changes in pectins of the Xylopodium of Ocimum nudicaule from dormancy to sprouting. Brazilian Journal Plant Physiology. 2006; 18: 325-331.
  209. 209. Alonso, AA, Machado, SR. Morphological and developmental investigations of the underground system of Erythroxylum species from Brazilian cerrado. Australian Journal of Botany. 2007: 55: 749-758.
  210. 210. Hoffman, WA, Schroeder W, Jackson, RB. Positive feedbacks of fire, climate, and vegetation and the conversion of tropical savanna. Geophysical Research Letters. 2002; 29:2052. doi:10.1029/2002GL015424
  211. 211. Smith, AM, Zeeman, SC. Annual Review of Plant Biology Starch: A Flexible, Adaptable Carbon Store Coupled to Plant Growth. Annual Review Plant Biology. 2020; 71: 9.1-9.29
  212. 212. Knapp, AK. Effect of fire and drought on the ecophysiology of Andropogon gerardii and Panicum virgatum in a tallgrass prairie. Ecology. 1985; 66: 1309-1320.
  213. 213. Furley, PA. The nature and diversity of neotropical savanna vegetation with particular reference to the Brazilian cerrados. Global Ecology and Biogeography. 2002; 8: 223-241.
  214. 214. Paula, SM, Naves, ER, Moreira, I, Padua, KJM, Silva, KR, Fernandes, WP. et al. Desempenho fotossintético de folhas jovens e maduras de Vochysia cinnamomea (Vochysiaceae) em áreas de cerrado rupestre intactas e pós-queimada. Bioiscience Journal 2015; 31(2): 591-600.
  215. 215. Graves, S.J, Rifai, SW, Putz, FE. Outer bark thickness decreases more with height on stems of fire-resistant than fire-sensitive Floridian oaks (Quercus spp.; Fagaceae). Americam Journal. Botany. 2014; 101: 2183-2188.
  216. 216. West, AG, Nel, JA, Bond, WJ, Midgley, JJ. Experimental evidence for heat plume-induced cavitation and xylem deformation as a mechanism of rapid post-fire tree mortality. New Phytologyst. 2016; 211: 828-838.
  217. 217. Simon, MF, Pennington, T. Evidence for fire adaptation to fire regimes in the tropical savannas of the Brazilian cerrado. International Journal Plant Science. 2012;173: 711-723.
  218. 218. Loram-Lourenço L, Farnese, FDS, Sousa, LFD, Alves, RDFB, Andrade, MCPD. et al. A Structure Shaped by Fire, but Also Water: Ecological Consequences of the Variability in Bark Properties Across 31 Species From the Brazilian Cerrado. Frontie Plant Science. 2020; 10:1718.
  219. 219. Martins, LC, Paiva, EAS. Flammable resin in Vellozia variabilis (Velloziaceae): Gland structure and chemical composition. Flora. 2016; 219: 94-100.
  220. 220. Rossatto, DR, Araújo, PE, Silva, BHP, Franco, AC. Photosynthetic responses of understory savanna plants: Implications for plant persistence in savannas under tree encroachment. Flora. 2017; https://doi.org/10.1016/j.flora.2017.12.009

Written By

Silvana Aparecida Barbosa de Castro and Vinícius Coelho Kuster

Submitted: 27 May 2020 Reviewed: 04 September 2020 Published: 08 October 2020