Open access peer-reviewed chapter

Mitigation of the Negative Impact of Warming on the Coffee Crop: The Role of Increased Air [CO2] and Management Strategies

Written By

Danielly Dubberstein, Weverton P. Rodrigues, José N. Semedo, Ana P. Rodrigues, Isabel P. Pais, António E. Leitão, Fábio L. Partelli, Eliemar Campostrini, Fernando Reboredo, Paula Scotti-Campos, Fernando C. Lidon, Ana I. Ribeiro-Barros, Fábio M. DaMatta and José C. Ramalho

Submitted: 27 June 2017 Reviewed: 10 November 2017 Published: 20 December 2017

DOI: 10.5772/intechopen.72374

From the Edited Volume

Climate Resilient Agriculture - Strategies and Perspectives

Edited by Ch Srinivasa Rao, Arun K. Shanker and Chitra Shanker

Chapter metrics overview

1,717 Chapter Downloads

View Full Metrics

Abstract

Crop sustainability can be threatened by new environmental challenges regarding predicted climate changes and global warming. Therefore, the study of real biological impacts of future environmental conditions (e.g., increased air [CO2], supra-optimal temperature and water scarcity) on crop plants, as well as the re-evaluation of management procedures and strategies, must be undertaken in order to improve crop adaptation and promote mitigation of negative environmental impacts, thus affording crop resilience. Coffee is a tropical crop that is grown in more than 80 countries, making it one of the world’s most traded agricultural products, while involving millions of people worldwide in the whole chain of value. It has been argued that this crop will be highly affected by climate changes, resulting in decreases in both suitable areas for cultivation and productivity, as well as impaired beverage quality in the near future. Here, we report recent findings regarding coffee species exposure to combined supra-optimal air temperatures and enhanced air [CO2], and impacts of drought stress on the crop. Ultimately, we discuss key strategies to improve coffee performance in the context of new environmental scenarios. The recent findings clearly show that high [CO2] has a positive impact on coffee plants, increasing their tolerance to high temperatures. This has been related to a better plant vigor, to the triggering of protective mechanisms, and to a higher functional status of the photosynthetic machinery. Even so, coffee plant is expected to suffer from water scarcity in a changing world. Therefore, discussion is focused on some important management strategies (e.g., shade systems, crop management and soil covering and terracing), which can be implemented to improve coffee performance and sustain coffee production in a continually changing environment.

Keywords

  • coffee crop sustainability
  • climate changes
  • mitigation
  • heat stress
  • drought

1. Introduction

Global emissions of the main greenhouses gases in the Earth’s atmosphere raised in the mid-eighteenth century during the industrial revolution associated with the use of fossil fuels. Since then, the CO2 concentration [CO2] has increased from 280 to 400 μL CO2 L−1 in 2014, and it is expected to rise to values between ca. 730 and 1020 μL CO2 L−1 by 2100 [1]. Agricultural activity has also directly contributed to this process, being responsible for 1/3 of the CO2 emissions but also with additional N2O and CH4 production, intensified mainly by inadequate management of crops and pastures [2], especially in low- and middle-income countries with predominating family farming [3].

Increased greenhouse gas emissions are expected to cause a temperature rise between 0.3 and 4.8°C by 2100, depending on future emissions and adequate measures to strongly limit them. Altered temperature may further promote extreme weather events, alter intra- and inter-annual precipitation patterns with long periods of drought and/or heavy rainfalls, partial melting of glacial ice, and consequently rising of the sea level [1]. Climate changes, particularly global warming, has a severe impact on the Earth’s ecosystem and pose serious threats to agricultural sustainability [4, 5, 6], which is one of the human activities most vulnerable to climatic variation, since plants require optimal growing conditions to produce desired quantity and quality products [7, 8]. On the other hand, global demand for food is increasing as it is linked to the rapidly growing populations, which together with climate constraints, may compromise world food security [9]. In addition, increase in [CO2] can affect the fundamental plant processes, such as photosynthesis and respiration, and, therefore, growth is also anticipated to be affected accordingly [10, 11, 12].

With regard to the coffee crop, it is known that plant growth, development, and productivity, as well as bean quality, are highly sensitive to climatic conditions [3, 13, 14, 15, 16]. Accordingly, recent modeling studies have predicted important reductions of suitable areas for coffee cultivation in several producing regions [7, 16, 17, 18, 19], with severe productivity losses in Mexico [20, 21], Nicaragua [3] and Tanzania [22], and extinction of wild populations of C. arabica in Ethiopia [23]. Although world coffee production has increased significantly in recent decades [24], studies state that climate change has caused substantial production losses [18], associated with periods of extreme droughts combined with supra-optimal temperatures [22, 25, 26], reducing coffee yields and bean quality as well as increasing the incidence of pests and diseases [16, 27]. In fact, it is believed that the recognized present climate changes have already caused yield losses in several coffee-producing countries, including Brazil, Ethiopia and Tanzania [22, 28, 29].

The negative estimates of future impacts on coffee crop were based on modeling approaches mostly focusing on increased air temperatures. However, these studies have only taken into account the current cultivars [30], and did not consider the considerable ability of some genotypes to endure various environment constraints, through metabolic adjustments and morphological and anatomical changes. Additionally, it was recently reported that coffee plants can respond positively to increased air [CO2] [31, 32, 33], improving plant physiological and metabolic performance, and mitigating warming impacts [11, 33, 34, 35]. Such beneficial effect could even overcome these impacts, allowing some yield increase under adequate water availability to the crop [36], particularly at higher altitudes [37]. Nevertheless, given that coffee is one of the most important agronomic products, and that possible implications of ongoing climate changes may affect the sustainability of this crop in many actual areas, with potential dramatic economic, social and environmental implications, there is an urgent need for improving our knowledge regarding the plant performance under a wide range of environmental conditions. It is also equally important to identify adequate mitigation and adaptation strategies to be implemented, such as shading system crop management and soil cover and terracing, together with breeding new cultivars, in order to alleviate the impacts of climate changes on coffee plants.

Studies dealing with water stress in coffee species and genotypes have provided a detailed picture of biological mechanisms involved in drought tolerance [38, 39, 40, 41, 42, 43, 44, 45, 46, 47, 48], whereas recent works also showed that some genotypes of both C. arabica and C. canephora can endure temperatures much higher than what was traditionally accepted [11, 35]. As referred, plant resilience can even be improved under the exposure to high atmospheric [CO2] [11, 34, 35, 36, 37, 49]. In this context, the objective of this review chapter is to report the recent findings regarding the coffee plant responses to the single and combined exposure to atmospheresupra-optimal temperatures and [CO2], as well as to drought stress, together with the envisagement of some important crop management strategies (e.g., intercropping/shade systems, soil covering and terracing), which can be implemented to improve coffee performance and to mitigate the impact of environmental constraints, aiming at sustaining coffee production in a permanent changing environment.

Advertisement

2. General aspects of production, origin and favorable environmental conditions for Coffea arabica and C. canephora

Coffee, one of the most traded commodities in the world, is supported by C. arabica L. and C. canephora Pierre ex A. Froehner species [14]. It is estimated that the coffee chain of value generates a global income of ca. US$ 173,000 million [50], having as well great social implications. In fact, this tropical crop is grown in approximately 80 countries [51], and about 25 million farmers, mainly smallholders, depend on this highly labor-intensive crop [52], with a worldwide involvement of ca. 125 million people in the entire chain [53]. Brazil, Vietnam, Colombia, Indonesia, Ethiopia, India, Honduras and Uganda are the major coffee producers, for a world annual production of green coffee beans which has been increasing steadily in the last decades, being consistently near or above 9 million tons since 2011/2012 [54]. This supports over 2.5 billion cups of coffee consumed every day around the world [55], with promising prospects for increased consumption in the coming years, especially among young people in Asia.

The coffee plant is characterized as a perennial woody shrub that belongs to the Rubiaceae family. Although there are at least 125 species within the Coffea genus [56], Coffea arabica L. (Arabica coffee) and Coffea canephora Pierre ex A. Froehner (Robusta coffee) are responsible for approximately 99% of the world coffee production [23, 57], with the former accounting for ~65% of total coffee production [55, 58]. Besides differences in origin, these species present important ecological differences in plant traits, as well in bean chemical composition, among them aroma precursors. In fact, the levels of these compounds have implications on sensory attributes, namely on astringency, taste, aroma, and flavor after roasting. Such chemical composition is not only genetic related but also strongly depend on environmental conditions (e.g., soil, shade, temperature), bean maturation stage, and to agricultural management and post-harvest procedures [59, 60, 61, 62, 63, 64].

C. arabica are originated from the tropical forests of Ethiopia, Sudan and Kenya, at altitudes of 1500–2800 m, annual averages air temperatures between 18 and 22°C, precipitation from 1600 to more than 2000 mm l distributed throughout the year, with a well-dry season (3–4 months), coinciding with the cold annual period. Currently, C. arabica coffee is grown in areas with cooler temperatures (18–23°C), at altitude mostly between 400 and 1200 m [7, 30, 65, 66], although cultivation up to 2000 m can be found in some countries in Central America. In contrast, C. canephora originated from the lowland forests of the Congo River basin, which extend to Lake Victoria in Uganda at altitudes up to 1200 m, are subjected to annual averages air temperatures between 23 and 26°C with minor fluctuations, and average precipitation exceeding 2000 mm distributed along 9–10 months [67, 68, 69]. Currently, cultivation occurs predominantly in lower altitude areas and higher temperatures, showing satisfactory development when the daily average temperature is above 22°C so that minimum is above 17°C and the average maximum air temperatures are below 31.5°C, with regular pattern of precipitation [70, 71, 72, 73, 74].

Advertisement

3. The impact of climate changes on coffee crop: warming and water scarcity

Coffee plants require both adequate water supply and optimal temperature, which are considered the most important environmental variables, since water and temperature-limited conditions cause negative impacts on growth, yield and productivity [14, 16, 30, 75]. Although in many coffee producing areas water scarcity occurs in the cooler season, climate modifications has increased the situations where low water availability and elevated temperature occur concomitantly under field conditions, which, as observed in other plants, will have the potential to exacerbate the limitations to the photosynthetic functioning [76].

In plants, photosynthesis and respiration are among the most sensitive metabolic processes to increasing temperatures [77]. High temperatures can cause protein denaturation and aggregation, increased production of reactive oxygen species [14], and ethylene synthesis [78]. Moreover, supra-optimal temperatures can reduce stomatal conductance and light energy use as well as alter thylakoid ultrastructure and diffusion of gas through mesophyll [15, 79, 80, 81] with a direct impact on net C gain. The latter will be even more amplified due to the increase of O2 solubility in relation to CO2 under higher temperatures, favoring the oxygenase activity of RuBisCO over its carboxylation activity, thus increasing photorespiration rates [82, 83]. Altogether, this ultimately may lead to the decline in the availability of carbohydrates for energy supply as well as carbon skeletons to support plant growth [77]. Thus, warmer temperatures can affect crop yield at any time from sowing to grain maturity, but it is the time around flowering, when the number of grains per land area is established, and during the grain-filling stage, when the average grain weight is determined, that high temperatures causes major impacts on the final harvestable crop [9, 73, 84]. In addition, it causes a reduction in the production of leaves and consequently alters the photosynthetic activity [85].

Coffee trees presented a remarkable tolerance to temperatures relatively high (up to 37/30°C; day/night) when air humidity was maintained at 75%, occurring relevant physiological/biochemical impairments only at 42/34°C, associated, namely, to large activity reductions of RuBisCO and Ru5PK [35], despite large accumulation of RuBisCO transcripts [86]. The reported heat tolerance was related with increases in protective molecules, namely, enzyme and non-enzyme antioxidant molecules, heat shock protein 70 (HSP70) reinforcement, and altered gene expression [11, 86]. However, under field conditions, rising temperature may lead to increase in air vapor pressure deficit (VPDair), what may result in decreased stomatal and canopy conductance in Coffea spp., due to a high sensitivity of coffee stomata to VPDair values above 2 kPa [87, 88, 89]. In addition, elevated temperatures can contribute to a gradual increase in soil water depletion, particularity in areas lacking sufficient precipitation, resulting in water stress, which further exacerbates the adverse effects of high temperatures.

Stomatal closure is one of the first responses to water deficit in coffee plants, aiming at limiting water loss through transpiration flow. However, this directly decreases the CO2 availability in the chloroplasts, reducing the photosynthetic rates [14]. In this context, irradiance reaching the chloroplasts may exceed the light energy needed to saturate photosynthesis, which in turn can lead to the formation of reactive oxygen species (ROS). ROS can cause oxidative damage to multiple cell and chloroplast components, namely to the D1 protein, lipids, RNA and DNA molecules, associated with increased cellular and metabolic disorders, resulting in cell death [47, 90, 91]. Moreover, ethylene synthesis often increases under drought stress conditions, promoting leaf senescence and slowing growth [10]. However, coffee plants display a noticeable metabolic plasticity to cope with environmental stresses [14, 51], as referred above for supra-optimal temperatures. Additionally, air [CO2] enrichment improved both coffee antioxidant defense system and photosynthetic performance regardless of temperatures, but maintaining a relevant photosynthetic functioning at temperature as high as 42°C. This prevented an energy overcharge in the photosynthetic apparatus, eventually reducing the need for energy dissipation and PSII photoinhibition [11, 35].

Considering water stress, a large number of early studies have reported that coffee plants can cope with drought stress through morphological, biochemical, and physiological modifications [14], as discussed later in this chapter. However, prolonged drought events associated with elevated temperatures can lead to very severe conditions, with a general impact on cell metabolism, associated as well to increased oxidative stress, altogether resulting in intense defoliation and yield losses (Figure 1), although genotypic difference in stomatal sensitivity to water stress among C. canephora genotypes have been reported [43, 45]. Furthermore, drought should be envisaged as contributing to a multidimensional stress, exacerbating the negative impacts of elevated irradiance and supra-optimal temperatures [13, 14, 42]. Therefore, drought-resistant coffee genotypes are a useful strategy for improving coffee performance in regions that are predicted to face moderate to severe drought [49].

Figure 1.

Intensive defoliation and yield losses due to prolonged severe drought (A), together with high temperatures (B) in C. canephora cultivations in Espírito Santo State, Brazil.

Overall, drought-sensitive C. canephora genotypes show a shallow root system and ineffective stomatal control, whereas drought-resistant coffee genotypes show considerably deeper root system, the strengthening of antioxidant defense system, and higher stomata sensitivity to reduced water availability (both in soil and atmosphere) [43, 92]. Increased wood density reinforcing vessels and, in turn improving resistance to cavitation, was correlated with tolerance to hydraulic dysfunctions [45]. On the other hand, C. canephora genotypes with specific traits conferring drought tolerance generally show reduced yield under optimal environments conditions due to their increased stomata sensitivity to VPDair. This is related to hydraulic limitations to water flow from roots to leaves [43, 45]. Therefore, coffee genotypes displaying increased phenotypic plasticity as, e.g., deep root system, substantial hydraulic conductance, intermediate stomatal control and strengthening of antioxidant defense system, could be used in regions which are predicted to face moderate water deficit, while drought-resistant genotypes could be used in regions predicted to face severe drought.

In addition to the traits outlined above, leaf size as well as canopy architecture should also be considered as important traits associated with drought tolerance. For example, although the leaf hydraulic conductivity (Kleaf) values found in C. arabica plants are typically low, probably linked to their native shade habitat [44, 93], C. arabica coffee genotypes with smaller leaves displayed higher vein density, higher Kleaf, increased gas exchange and reduced drought vulnerability [40, 44]. Drought tolerance was also found to be higher for C. canephora genotypes displaying smaller leaves [42]. In fact, it is known in other plants that smaller leaves allow for more rapid convective heat loss, resulting in lower transpiration and water loss likely due to smaller boundary layer [94]. Furthermore, a more compact crown structure may result in reduced VPDair within the coffee canopy, decreasing the transpiration demand [14], besides allowing to increase plant density coupled with improved soil covering and reducing the negative impacts of elevated temperatures, and high wind speed on coffee trees. On the other hand, C. arabica genotypes displaying open architecture crown show high transpiration rates (as measured by the sap flow technique) depleting accessible soil water more rapidly [40]. Therefore, although the water use efficiency in coffee genotypes is associated with the hydraulic capacity of the soil and stem to supply the leaves with water [95], coffee traits linked to water safety, e.g., a more compact crown structure and to greater extent an effective stomatal control, seem to play an important role in drought tolerance.

A recent study by [48] reported that both drought-sensitive and drought-tolerant C. canephora genotypes showed a drought stress “memory,” with plants exposed to multiple drought events showing better recovery than those submitted to drought events for the first time. This performance was mainly associated with substantial metabolic reprogramming, involving key processes such as photosynthesis, respiration, photorespiration, and the antioxidant system. In this sense, it would appear reasonable to suggest that multiple moderate water stress in coffee seedlings at nursery stage may improve to some extent the initial coffee performance under field conditions in areas prone to water scarcity.

Advertisement

4. Can elevated [CO2] help the mitigation of the negative impacts of high temperature and water deficit?

Although climate models point CO2 as the major greenhouse gas responsible for global warming due to its high accumulation rate in the atmosphere [6], the impacts of increased air [CO2] at plant physiological and biochemical levels should not be neglected, namely in coffee metabolism [11, 31, 32, 35], as well in yield [36, 37].

The current [CO2] in the atmosphere is still below the optimum for photosynthesis of C3 plants; therefore, leaf photosynthetic rates are predicted to increase in response to future increase in air [CO2], due to increased carboxylase activity of RuBisCO [82, 83, 96]. This C-fertilization may eventually reinforce plant vigor (and the defense systems), which, in turn, could reinforce the plant ability to endure environmental stresses [97]. On the other hand, elevated CO2 levels will especially benefit plants with strong sink capacity to use such increased amounts of photoassimilates. Otherwise, an accumulation of soluble sugars may occur which in turn will decrease the net photosynthetic rate through negative feedback mechanisms, that is, will provoke downregulation of photosynthesis, not allowing the plant to fully explore the positive effect of [CO2] increase [83].

In the case of coffee, significant increases of net photosynthesis, between 34 and 49%, were observed for C. canephora (Clone 153) and C. arabica (Icatu and IPR 108) genotypes [31], when comparing plants grown subjected to elevated [CO2] (700 μL L−1) or normal [CO2] (380 μL L−1) under environmental controlled conditions. Furthermore, under such high [CO2], plants also showed a better water-use efficiency, reinforcement of photosynthetic components and increased activity of key enzymes involved in photosynthesis and respiration, without noticeable leaf sugar accumulation. Therefore, these coffee genotypes were able to cope with enhanced [CO2], maintaining the consumption of photosynthates and regeneration of RuBP associated with continuous investment in vegetative and reproductive structures. The evidence of improved coffee performance under enhanced [CO2] was further obtained with other C. arabica genotypes (Obatã IAC 1669–20 and Catuaí Vermelho IAC 144) under field conditions using free-air CO2 enrichment (FACE) system, showing increased photosynthesis and decreased photorespiration, without changes in stomatal and mesophyll conductance, for an air [CO2] of 550 μL L−1 [33]. Additionally, coffee plants grown under elevated [CO2] were more vigorous, with increased leaf area, growth rate at height and stem diameter, showing as well increased grain yield by 14.6 and 12.0% for Catuaí Vermelho 144 and Obatã IAC 1669–20, respectively, [8, 32], although average yield increases of 28% were also reported after three harvests [37] when compared to plant grown at ambient [CO2]. Another study also demonstrated that coffee trees grown under 550 μL CO2 L−1 presented increase in photosynthesis of leaves from upper and lower canopy layers, inhibition of photorespiration, and no apparent sign of photosynthetic downregulation, when compared to plants grown under ambient [CO2] (390 μL L−1) [98]. Finally, recent studies based on modeling approaches accounting with high air [CO2] positive impact reported that coffee yield losses associated mostly with high temperatures can be offset by the CO2 fertilization effect, with a probably yield increase by 2040–2070 [36], or 2050, particularly at higher altitudes [37].

The simultaneous occurrence of various environmental constraints is the most common situation under field conditions, and therefore, it has been argued that a positive plastic response from plant experiencing a single stress can be increased, canceled or even reverted under the combined action of multiple stresses [6]. Regarding the coffee plant, responses to the combined effects of increased [CO2] and supra-optimal air temperature started to be investigated quite recently, whereas the simultaneous exposure to elevated [CO2], heat and water deficit have never been studied. The exposure to increased air [CO2] revealed interesting implications to plant physiological response to supra-optimal conditions. This was the case in both C. arabica (cvs. Icatu and IPR 108) and C. canephora cv. Conilon Clone 153 plants exposed to elevated [CO2] and temperatures up to 42°C [11, 34, 35]. Notably, a remarkable heat tolerance was observed up to 37/30°C (day/night) irrespective of air [CO2]. The tolerance (and high physiological performance) to such temperature was somewhat surprising as it is above what is traditionally accepted to be tolerated by coffee plant [35]. Furthermore, enhanced [CO2] greatly mitigated the negative impact of the temperature, especially at 42/34°C, with higher water-use efficiency (WUE) at moderately higher temperature (31/25°C). Increased CO2 was observed to strengthen the photosynthetic apparatus, improving light energy use and biochemical functioning. These results were linked to the maintenance or increase in the content of several protective molecules (neoxanthin, lutein, β-carotene, α-tocopherol, heat shock protein-HSP70, raffinose), the activity of antioxidant enzymes (superoxide dismutase, SOD; ascorbate peroxidase, APX, glutathione reductase, GR; catalase, CAT) and the upregulation of some genes related to stress-protective molecules (ELIP, HSP70, Chaperonin 20 and 60), and antioxidant enzymes (CAT, CuSOD2, APX Cyt, APX Chl) [11]. In the same experiments, overall leaf mineral macro- and microelement contents have remained within a range that could be considered largely adequate for coffee plants, with no changes in macronutrient profile (N > K > Ca > Mg > S > P), that is, satisfactory mineral content was maintained in the context of warming, under high [CO2] [34].

Climate changes are also predicted to affect intra- and inter-annual rainfall patterns, and the decrease in precipitation amounts in conjunction with increased air temperature may reduce net photosynthesis at current [CO2]. Still, under increased air [CO2], a partial relief of negative impacts of water deficit may occur [99]. Indeed, arabica coffee plants grown under severe drought conditions and increased biotic pressure showed strategies which allow the maintenance of structural and physiological integrity in the fourth period of winter growth [98]. This occurs because of the dichotomous responses of net photosynthesis and stomatal conductance to high [CO2], which lead to improved WUE, reducing soil moisture depletion during periods of drought [9]. Studies by [10, 76] on Agropyron cristatum L. and Perilla frutescens var. japonica Hara, respectively, reported positive results of elevated-CO2 mitigation of drought stress, verifying increase in photosynthetic capacity and decrease in stomatal conductance with lower transpiration rates. Consequently, increased intrinsic water-use efficiency (WUEi) and total water-use efficiency (WUEt) were observed. Furthermore, high [CO2] can also alleviate oxidative stress conditions, and photoinhibition status, likely associated to a higher photosynthetic functioning (as also observed for high temperatures [11]), even under significant stomatal closure. Altogether such responses may result in improved tolerance to drought stress, as found in other plants [6, 10, 12]. Nevertheless, it is important to note that under severe drought, such positive results might not be obtained, and that mitigation associated with high [CO2] does not always occur [6].

In addition to the positive effects on the impacts of abiotic stresses, elevated [CO2] can also reduce to some extent the incidence and severity of coffee pests and diseases. In fact, decrease in leaf rust (Hemileia vastatrix) severity, number of lesions, leaf area injured, number of sporulating lesions, percentage of damaged leaf area and area under disease progress were observed in C. arabica cv. Catuaí IAC 144 grown under elevated [CO2] [8]. Reduced incidence of leaf miner (Leucoptera coffeella) during periods of high infestation was also observed at elevated [CO2] [32].

In summary, enhanced [CO2] can have a positive mitigation effect on the negative impacts of high temperature and, probably, low water availability, as well as by reducing the severity of some pests and diseases. However, since responses are highly species (and even cultivar) dependent, it is urgent to implement long-term studies in coffee considering single and, especially, combined stresses, with the simultaneous exposure to elevated [CO2], supra-optimal temperatures and drought, relating them to phenological stages (e.g., flowering), therefore, to increase knowledge on this crop in a context of climate changes.

Advertisement

5. Mitigating the impacts of climate changes through management practices

To promote crop sustainability in the context of climate changes and global warming, adaptation and mitigation measures must be implemented. Regarding adaptation, plant screening and breeding are essential to provide new improved and stress-tolerant genotypes, but their implementation are somewhat delayed due to the time needed to obtain new varieties. As an example, the use of improved genotypes with an optimized architecture is a valuable tool. It is known that small-size plants, with denser canopies, are prone to display lower transpiration rates [13, 14]. Additionally, plants with larger and deeper root systems would have an ability to explore increased soil volumes, reaching water resources that other plants with a more superficial root system do not [14]. Still, several years will be needed until such new genotypes can be available and, therefore, ready-to-use strategies should be implemented, namely those regarding an effective mitigation of the environmental negative impacts on the actual cropped genotypes. This can be even more important when dealing with tree crops that have a productive lifespan of several years or decades, as it is the case of coffee, which can last for more than 30 years [18].

A significant range of management techniques can be used to minimize the impact of different stresses that can affect the performance of agricultural systems. For coffee crop, several different agronomic tools stand to that purpose, e.g., the use of shade systems with tree species, as well as other intercropping associations, to improve an efficient water use and minimize warming at the plant level, maintaining a more suitable microenvironment concerning both temperature and air humidity. Improved soil covering with other intercropped species, and terracing under conditions of significant slopes, are also quite useful techniques to minimize soil water loss (or to increase its infiltration), therefore, helping to maintain water resources available to the plants for longer periods.

5.1. Fertilization management under high air [CO2] and warming conditions

Minerals have a wide number of roles in plant cell. Therefore, as in other plants, an adequate mineral fertilization is recognized as crucial to allow the triggering of acclimation mechanisms in face of environmental constraints in the coffee plant. This is the case of nitrogen (N) supply, which is of utmost importance to allow the recovery from high irradiance impact, through the triggering of repair mechanism, and the reinforcement of leaf defense mechanisms, including the control of highly reactive molecules of chlorophyll and oxygen, whose production is exacerbated under high irradiance/full sun exposure [100, 101, 102]. Additionally, the presence of adequate contents of other minerals allows the plant to maintain high metabolic performance due to their specific roles. For instance, copper, iron and manganese, which were shown to promote the activities of, respectively, superoxide dismutase, ascorbate peroxidase, and photosystem II under cold exposure [103], as well as calcium, which is essential to the stabilization of chlorophyll and the maintenance of photochemical efficiency at PS II level [104].

Changes in mineral contents may affect plant development, but may also have other important consequences, namely as regards the quality of agricultural products for food and feed, herbivory, litter decomposition rates, etc. [105, 106]. It is known that mineral contents often decline in the leaf biomass under high air [CO2] conditions. This was related to higher growth rates, accumulation of non-structural sugars (mainly starch), lower transpiration rates, or to changes in the nutrient allocation patterns under enhanced air [CO2] [107, 108, 109] This mineral “dilution” effect on leaves can affect the photosynthetic apparatus (e.g., through N, S and Fe), enzyme activity (e.g., through K, P, Mn and Fe), alters redox reactions (e.g., through Fe, Zn and Cu), and modifies the structural integrity of chloroplast membranes (e.g., B) [105, 110, 111, 112, 113]. However, this so called “dilution effect” may frequently reflect qualitative physiological changes rather than a lack of nutrients [108], since in many cases, these plants did not present mineral nutrition disturbances. This seems just to be the case in Coffea spp., since it was observed that under adequate temperature, long-term exposure to enhanced [CO2] (700 μL L−1) net photosynthetic rate was increased by between 40 and 49% [31], concomitantly to a moderate mineral reduction that ranged from 7 to 25% in N, Mg, Ca, Fe in C. canephora cv Conilon Clone 153, and in N, K and Fe in C. arabica cv. Icatu [34].

Most important was also the observation that contents (on a per leaf mass basis) of several minerals increased under supra-optimal temperatures, largely offsetting the dilution effect observed under control temperature (25°C), keeping the large majority of minerals and their ratios within a range that is considered adequate, therefore, suggesting that coffee plant can maintain its mineral balance in a context of climate changes and global warming [34]. Even so, taking into account the importance of mineral dynamics to virtually all biological processes, studies under field conditions must be implemented to better understand the possible CO2 implications for coffee fertilizer management in a context of climate changes and global warming in a near future.

5.2. Reducing irradiance at the leaf level

Both C arabica and C. canephora have been cultivated under full sunlight in many regions around the world, particularly in Brazil. In fact, coffee plant can successfully adjust its photosynthetic metabolism to high light conditions, namely if adequate mineral nutrition is provided [100, 101, 102]. Effective acclimation to other environmental constraints (e.g., cold, heat, drought) was also reported [14]. Such acclimation ability depends on the presence and/or reinforcement of several mechanisms, among them leaf antioxidants, and qualitative modifications on the lipid matrix of cell membranes, particularly in the chloroplast. This allows the plant to maintain high metabolic activity, namely as regards the photosynthetic pathway, depending on stress severity and on species and genotype capabilities [11, 41, 57, 93, 101, 114]. However, these coffee species have evolved and grow naturally under shaded understory [14, 68, 69]. Not surprisingly, Coffee sp. presents some leaf traits usually associated with shade plants, namely low light saturating point (ca. 500 μmol m−2 s−1) [115], therefore, quite below the irradiance values occurring under field conditions. This increases the probability of photoinhibition under high solar radiation [13, 14, 100, 116, 117]. Taking into account predictions of a global warming and lower water availability along the present century, the implementation of coffee cultivation under shaded conditions (e.g., under agro-forestry systems) may be recommended as a cultural management practice to alleviate the combined impacts of drought and elevated temperatures [118], while improving nutrient cycling, soil fertility and soil organic matter accumulation [119, 120, 121, 122]. Additionally, shade crops can improve ecological aspects including increasing bio-diversity of flora and fauna [123, 124].

Traditionally, coffee trees grown under shaded conditions show reduced yield, since shade trees may compete with coffee for essential requirements such as light, water and nutrient depending on tree density [13, 119, 125], with less nodes per branch and fewer flowers at existing nodes must be also considered. Additionally, coffee plants show limited light distribution within their own canopies [88], thus leading to the further reduction of the light availability at whole canopy scale. However, increased light-use efficiency can compensate the low availability of photosynthetically active solar radiation in coffee trees grown under shaded conditions [126]. Also, shade trees can increase the proportion of diffuse light under their canopy by 60–90%, what may lead to increased penetration of radiation inside the coffee canopy [126]. In fact, C. canephora Clone 02 (clonal variety “EMCAPA 8111” [127]), grown under an irradiance retention of 70% promoted by Australian cedar (Toona ciliata M. Roem) in southeastern region of Brazil showed similar yield to unshaded counterparts, although for a study considering only one crop season [128] (Figure 2). Similar yield and leaf nutrient content were also found in shaded C. canephora cv. Verdebras G35 plants intercropped with rubber trees (Hevea brasiliensis (Willd. ex A. Juss.) Müll. Arg.) in the same region, with a reduction of ca. 70% in total irradiation [129], while similar yield were reported for C. arabica cv. Caturra intercropped with Erythina poeppigiana (reduction of ca. 70% in total irradiation) in the central Valley of Costa Rica [126] and in six C. arabica genotypes shaded by E. verna and Musa sp. (shade up to 80%) [130].

Figure 2.

Coffea canephora cv. Conilon under shading conditions promoted by A) Australian cedar (Toona ciliate M. Roem. var. Australis), B) papaya (Carica papaya L.), C) rubber tree (Hevea brasiliensis Willd. ex A. Juss), and D) African mahogany (Khaya spp.), in northern Espírito Santo state, Brazil.

As referred above, coffee trees show increased stomatal sensitivity to VPDair, so that increase in air temperature and/or decrease in air relative humidity (RH) can result in reduced stomatal aperture. In this sense, shaded systems with trees, including rubber [129] and Australian cedar [128], can reduce air temperature, maintain higher air humidity, and decrease low wind speed near the coffee plants, thus resulting in decreased VPDair between the leaf and the atmosphere, and a lower water loss through transpiration [13]. Therefore, shade will promote a better WUE, reducing plant transpiration and soil evaporation, while contributing to improve plant physiological performance [117].

In addition to the impacts on photosynthetic machinery, rising temperature causes increases in plant respiration rates, mainly associated with “maintenance respiration” to support protein turnover and to maintain active ions transport across the membrane [81]. Recent studies have reported decreases of 2 up to 6°C in air temperature surrounding coffee canopy under shaded condition [125, 128, 129, 131]. Such reduction in air temperature can therefore reduce maintenance respiration [126], as C. arabica cv. Caturra plants in the Central Valley of Costa Rica that showed a 40% decrease in peak maintenance respiration under a 4°C decrease in maximum temperature [125].

Coffee growers need to obtain high yields, while maintaining bean quality in order to guarantee their income. Rising temperature may decrease coffee bean yields due to bud abortion or development of infertile flowers, particularly when associated with prolonged dry periods [65]. Additionally, increased temperature may accelerate fruit maturation and ripening, reducing the accumulation of sucrose and altering the content of several compounds that are known precursors of taste, flavor and aroma after roasting [15, 60, 62, 64]. Shade trees may provide a milder microclimate, attenuating temperature rise on coffee beans, and by lowering air temperature close to the coffee plant can extend the maturation period so that the bean filling period will be enlarged [132, 133], what can contribute to higher sucrose accumulation.

Besides the importance of shade in reducing thermal stress, other important benefits arise as well. For instance, coffee trees grown under full sunlight show a typical biennial pattern, e.g., during one crop season, a heavy fruit load will constitute a major sink at the expense of new leaves and branches, reducing productivity in the following year [134]. Moreover, high fruits load may result in reduced bean size due to the carbohydrate competition among berries during bean filling [133]. In this sense, depending on density, shade trees can reduce coffee flowering intensity, resulting in a better coffee bean quality, as well as in higher yield stability along the years. Although the central purpose of coffee cultivation under shaded conditions is alleviating the impacts of both high irradiances and supra-optimal temperatures, it is worth to mention that cultivation of trees of economic importance, such as Inga sp. [125], Australian cedar [128], rubber tree (Figure 2) [129], can constitute important complementary sources of income to coffee farmers.

The application of kaolin particles can also reduce the irradiance at leaf surface, increasing radiation reflections, and, consequently decreasing leaf temperature [135]. Kaolin particle film can as well improve light distribution inside the canopy, leading to increase in photosynthetic rates, increasing crop water use efficiency at whole-canopy scale, as reported for apple (Malus sylvestris) [136, 137] and grapevine (Vitis vinifera L.) [137]. Moreover, kaolin particle film protected apple fruits from damage caused by excessive heat linked to high light conditions, besides avoiding the direct impacts of ultraviolet radiation on fruits as well [135]. Additionally, some works have demonstrated that particle film technology can alleviate the negative impacts of water stress, particularly associated with increase in light reflection and decrease in canopy temperature [137, 138]. In coffee, kaolin particle film was observed to increase C-assimilation and bean yield, linked to improved light distribution within the canopy, since sunlight is essential to floral initiation [139], and can, therefore, constitute a promising alternative technique to reduce the thermal energy at leaf level.

Considering the effects of supra-optimal temperatures, high density planting system can alleviate the negative impacts of heat stress, because under such conditions, the air surrounding the coffee plants becomes more humid due to plant transpiration and low wind speed, decreasing VPDair [14]. Additionally, in areas facing strong winds, the use of windbreaks or tree shelters is recommended as both can avoid an extensive removal of boundary layer, leading to decreased demand for water from the atmosphere. However, under high density planting systems, coffee crop management through pruning is fundamental for renewal, revitalizing and yield stability in coffee plantations [140], what can improve soil coverage.

5.3. Soil covering and terracing

The distance between coffee rows allows for growth of other plants, which may compete for water and nutrients, depending on species involved. Overall, weed control aims at removing the invasive plants, exposing soil to intense solar radiation which can result in increase in water evaporation directly from the soil as well as facilitating the surface water runoff, leading to erosion losses, especially in areas with a pronounced slope. Depending on weed species, invasive plants are allowed to grow naturally between coffee rows without any management strategy. Although such plants may reduce erosion losses and direct solar radiation, as well as improve the infiltration of water into the soil stratum [141], they lose water during the day through transpiration, decreasing soil moisture [142]. Therefore, weed management strategies (for example, cut using a mower) can contribute for organic matter accumulation and, in turn, increase the water retention capacity of the soil, improving water productivity.

Also, the use of some leguminous species, correctly managed between coffee rows, can protect the soil, providing N to the coffee plants. Furthermore, soil coverage with herbaceous plants between coffee rows increases soil moisture and reduces both soil temperature and weed incidence, improves the physical and chemical soil properties [143, 144], promotes water infiltration, reduces rainfall impact and erosion, stimulates microbial activity, and improves organic matter in the soil [145]. Improved ground cover can be further obtained from weeds control, and by keeping biomass from coffee plants pruning, a common practice used to promote crop productivity [140] and soil microbiota diversity.

Coffee straw/husks, a by-product generated during coffee processing and discarded in many farms, can also be used for soil covering, reducing water losses through soil evaporation. In addition, coffee straw/husks can provide essential macro and micronutrients, namely N, P, K, Ca, Mg, S, Fe, B, Mn, Zn and Cu [72], lowering the need of chemical fertilization regarding these nutrients, and increasing coffee yield up to 25% [146]. Moreover, these coffee by-products can improve the soil physical associated with increase in CTC and soil pH [147], and inhibit seed germination of many weed species such as Amaranthus retroflexus, Bidens pilosa, Cenchrus echinatus and Amaranthus spinosus [148]. Therefore, coffee straw/husks can increase soil water retention and reduce to some extent costs associated with weed managements and fertilizers.

Other strategies for areas with a high slope are terracing, contour plowing terrace and rectangular ditches. Such practices contribute for preventing rapid surface runoff, allowing rain water to percolate into the soil, contributing for soil conservation [149, 150, 151]. Therefore, the establishment of terraces, although expensive, could constitute a worthwhile alternative to reduce water losses through runoff and soil erosion, while promoting infiltration [152]. Rain water storage in reservoirs should also be implemented. This will allow future water use during periods of negligible rainfall, constituting an important mitigation strategy to avoid drought stress. Therefore, increasing the water retention/storage capability in the farm can delay or even prevent coffee water stress.

Advertisement

6. Future perspectives

Climate changes are expected to negatively affect the coffee crop, causing serious social and economic impacts. Supra-optimal temperatures and water scarcity may decrease coffee yields and some studies state that these stresses are already occurring in some coffee-growing countries. However, coffee plants show a potential ability to cope with several environmental stresses and enhanced [CO2] can improve such ability and mitigate to some extent the negative impacts of supra-optimal temperatures. Even so, some mitigation strategies will be necessary to alleviate the impacts of elevated temperature and/or drought stress on coffee trees. We have reviewed some strategies that can be implemented depending on main environmental stresses occurring in specific regions, such as those based on coffee traits (root systems, size leaf, canopy architecture and stomatal sensitivity) and crop management (nutrient managements and pruning system), as well as those aiming at reducing excessive light at coffee tree level (shaded systems, kaolin-based particle film and plant density), and at improving soil water retention (soil covering and terracing). Notably, however, a single mitigation strategy may not be enough to face severe stress conditions; thus, multiple strategies should be undertaken.

Future studies considering simultaneous exposure to the main environmental stresses (e.g., high temperatures and drought), taking into account as well elevated [CO2], will be necessary to elucidate the mechanisms underlying plasticity and vulnerability of coffee plants under conditions that are expected to occur in the fields in a near future. Such studies are a fundamental basis for plant breeders to obtain new/more adapted genotypes. Finally, these strategies appear to be useful tools toward maintaining the coffee chain production.

Advertisement

Acknowledgments

This work was partly supported by contribution of national funds from Fundação para a Ciência e a Tecnologia through the research units UID/AGR/04129/2013 (LEAF) and UID/GEO/04035/2013 (GeoBioTec). Brazilian funding from CAPES (grant PDSE: 88881.132375/2016-01, Danielly Dubberstein), CNPq and Fapemig (fellowships to F.M. DaMatta, F. Partelli, E. Campostrini and W.P. Rodrigues-PDJ438817/2016-8) are also greatly acknowledged.

References

  1. 1. IPCC. Climate change 2014: Mitigation of climate change. In: Edenhofer O, Pichs-Madruga R, Sokona Y, Farahani E, Kadner S, Seyboth K, Adler A, Baum I, Brunner S, Eickemeier P, Kriemann B, Savolainen J, Schlömer S, von Stechow C, Zwickel T, Minx JC, editors. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge, United Kingdom and New York, NY, USA: Cambridge University Press; 2014
  2. 2. Van Beek CL, Meerburg BG, Schils RLM, Verhagen J, Kuikman PJ. Feeding the world’s increasing population while limiting climate change impacts: Linking N2O and CH4 emissions from agriculture to population growth. Environmental Science & Policy. 2010;13:89-96. DOI: https://doi.org/10.1016/j.envsci.2009.11.001
  3. 3. Rahn E, Läderach P, Baca M, Cressy C, Schroth G, Malin D, van Rikxoort H, Shriver J. Climate change adaptation, mitigation and livelihood benefits in coffee production: Where are the synergies? Mitigation and Adaptation Strategies for Global Change. 2013;19:1119-1137. DOI: 10.1007/s11027-013-9467-x
  4. 4. Beach RH, Cai Y, Thomson A, Zhang X, Jones R, Mccarl BA, Crimmins A, Martinich J, Cole J, Ohrel S, Deangelo B, Mcfarland J, Strzepek K, Boehlert B. Climate change impacts on us agriculture and forestry: Benefits of global climate stabilization. Environmental Research Letters. 2015;10:095004. DOI: 10.1088/1748-9326/10/9/095004
  5. 5. Tack J, Barkley A, Nalley LL. Effect of warming temperatures on US wheat yields. Proceedings of the National Academy of Sciences of the United States of America. 2015;112:6931-6936. DOI: 10.1073/pnas.1415181112
  6. 6. Xu Z, Shimizu H, Yagasaki Y, Ito S, Zheng Y, Zhou G. Interactive effects of elevated CO2, drought, and warming on plants. Journal of Plant Growth Regulation. 2013;32:692-707. DOI: 10.1007/s00344-013-9337-5
  7. 7. Zullo Jr J, Pinto HS, Assad ED, Ávila AMH. Potential for growing Arabica coffee in the extreme south of Brazil in a warmer world. Climatic Change. 2011;109:535-548. DOI: 10.1007/s10584-011-0058-0
  8. 8. Tozzi FRO, Ghini R. Impact of increasing atmospheric concentration of carbon dioxide on coffee leaf rust and coffee plant growth. Pesquisa Agropecuária Brasileira. 2016;51:933-941. DOI: 10.1590/S0100-204X2016000800005
  9. 9. DaMatta FM, Grandis A, Arenque BC, Buckerigde MS. Impacts of climate changes on crop physiology and food quality. Food Research International. 2010;43:1414-1423. DOI: https://doi.org/10.1016/j.foodres.2009.11.001
  10. 10. Lee SH, Woo SY, Je SM. Effects of elevated CO2 and water stress on physiological responses of Perilla frutescens Var. japonica HARA. Plant Growth Regulation. 2015;75:427-434. DOI: 10.1007/s10725-014-0003-0
  11. 11. Martins MQ, Rodrigues WP, Fortunato AS, Leitão AE, Rodrigues AP, Pais IP, Martins LD, Silva MJ, Reboredo FH, Partelli FL, Campostrini E, Tomaz MA, Scotti-Campos P, Ribeiro-Barros AI, Lidon FJC, DaMatta FM, Ramalho JC. Protective response mechanisms to heat stress in interaction with high [CO2] conditions in Coffea spp. Frontiers in Plant Science. 2016;29:947-964. DOI: https://doi.org/10.3389/fpls.2016.00947
  12. 12. AbdElgawad H, Zinta G, Beemste GTS, Janssens IA, Asard H. Future climate CO2 levels mitigate stress impact on plants: Increased defense or decreased dhallenge? Frontiers in Plant Science. 2016;7:556-562. DOI: https://doi.org/10.3389/fpls.2016.00556
  13. 13. DaMatta FM. Ecophysiological constraints on the production of shaded and unshaded coffee: A review. Field Crops Research. 2004;86:99-114. DOI: https://doi.org/10.1016/j.fcr.2003.09.001
  14. 14. DaMatta FM, Ramalho JC. Impacts of drought and temperature stress on coffee physiology and production: A review. Brazilian Journal of Plant Physiology. 2006;18:55-81. DOI: http://dx.doi.org/10.1590/S1677-04202006000100006
  15. 15. Santos CAFD, Leitão AE, Pais IP, Lidon FC, Ramalho JC. Perspectives on the potential impacts of climate changes on coffee plant and bean quality. Emirates Journal of Food and Agriculture. 2015;27:152-163. DOI: 10.9755/ejfa.v27i2.19468
  16. 16. Ovalle-Rivera O, Läderach P, Bunn C, Obersteiner M, Schroth G. Projected shifts in Coffea arabica suitability among major global producing regions due to climate change. PLoS One. 2015;10:e0124155. DOI: 10.1371/journal.pone.0124155
  17. 17. Zullo Jr J, Pinto HS, Assad ED. Impact assessment study of climate change on agricultural zoning. Meteorological Applications. 2006;13:69-80. DOI: 10.1017/S135048270600257X
  18. 18. Bunn C, Läderach P, Rivera OO, Kirschke D. A bitter cup: Climate change profile of global production of Arabica and Robusta coffee. Climatic Change. 2015;129:89-101. DOI: 10.1007/s10584-014-1306-x
  19. 19. Magrach A, Ghazoul J. Climate and pest-driven geographic shifts in global coffee production: Implications for forest cover, biodiversity and carbon storage. PLoS One. 2015;10:e0133071. DOI: https://doi.org/10.1371/journal.pone.0133071
  20. 20. Gay C, Estrada F, Conde C, Eakin H, Villers L. Potential impacts of climate change on agriculture: A case of study of coffee production in Veracruz, Mexico. Climatic Change. 2006;79:259-288. DOI: 10.1007/s10584-006-9066-x
  21. 21. Schroth G, Laderach P, Dempewolf J, Philpott S, Haggar J, Eakin H, Castillejos T, Moreno JG, Pinto LS, Hernandez R, Eitzinger A, Ramirez-Villegas J. Towards a climate change adaptation strategy for coffee communities and ecosystems in the sierra Madre de Chiapas, Mexico. Mitigation and Adaptation Strategies for Global Change. 2009;14:605-625. DOI: 10.1007/s11027-009-9186-5
  22. 22. Craparo ACW, Van Asten PJA, Läderach P, Jassogne LTP, Grab SW. Coffea arabica yields decline in Tanzania due to climate change: Global implications. Agricultural and Forest Meteorology. 2015;207:1-10. DOI: https://doi.org/10.1016/j.agrformet.2015.03.005
  23. 23. Davis AP, Gole TW, Baena S, Moat J. The impact of climate change on indigenous arabica coffee (Coffea arabica): Predicting future trends and identifying priorities. PLoS One. 2012;7:e47981. DOI: https://doi.org/10.1371/journal.pone.0047981
  24. 24. Jha S, Bacon CM, Philpott SM, Méndez VE, Läderach P, Rice RA. Shade coffee: Update on a disappearing refuge for biodiversity. Bioscience. 2014;64:416-428. DOI: https://doi.org/10.1093/biosci/biu038
  25. 25. Pezzopane JRM, Castro FS, Pezzopane JEM, Bonomo R, Saraiva GS. Climatic risk zoning for Conilon coffee in Espirito Santo, Brazil. Revista Ciência Agronômica. 2010;41:341-348. DOI: 10.1590/S1806-66902010000300004
  26. 26. van der Vossen H, Bertrand B, Charrier A. Next generation variety development for sustainable production of arabica coffee (Coffea arabica L.): A review. Euphytica. 2015;204:243-256. DOI: 10.1007/s10681-015-1398-z
  27. 27. Villers L, Arizpe N, Orellana R, Conde C, Hernandez J. Impacts of climatic change on coffee flowering and fruit development in Veracruz, México. Interciencia. 2009;34:322-329. DOI: 0378-1844/09/05/322-08 $ 3.00/0
  28. 28. Conab. Acompanhamento da safra brasileira de café, v. 3 – Safra 2016, n. 4 – Quarto Levantamento, Brasília. 2016. pp. 1-77
  29. 29. Moat J, Williams J, Baena S, Wilkinson T, Gole TW, Challa ZK, Demissew S, Davis AP. Resilience potential of the Ethiopian coffee sector under climate change. Nature Plants. 2017;3:art17081. DOI: 10.1038/nplants.2017.81
  30. 30. Camargo MBP. The impact of climatic variability and climate change on arabica coffee crop in Brazil. Bragantia. 2010;69:239-247. DOI: http://dx.doi.org/10.1590/S0006-87052010000100030
  31. 31. Ramalho JC, Rodrigues AP, Semedo JN, Pais IP, Martins LD, Simões-Costa MC, Leitão AE, Fortunato AS, Batista-Santos P, Palos IM, Tomaz MA, Scotti-Campos P, Lidon FC, DaMatta FM. Sustained photosynthetic performance of Coffea spp. under long-term enhanced [CO2]. PLoS One. 2013;8:e82712. DOI: 10.1371/journal.pone.0082712
  32. 32. Ghini R, Torre-Neto A, Dentzien AFM, Gerreiro-Filho O, Iost R, Patrício FRA, Prado JSM, Thomaziello RA, Bettiol W, Damatta FM. Coffee growth, pest and yield responses to free-air CO2 enrichment. Climatic Change. 2015;132:307-320. DOI: 10.1007/s10584-015-1422-2
  33. 33. DaMatta FM, Godoy AG, Menezes-Silva PE, Martins SCV, Sanglard LMVP, Morais LE, Torre-Neto A, Ghini R. Sustained enhancement of photosynthesis in coffee trees grown under free-air CO2 enrichment conditions: Disentangling the contributions of stomatal, mesophyll, and biochemical limitations. Journal of Experimental Botany. 2016;67:341-352. DOI: 10.1093/jxb/erv463
  34. 34. Martins LD, Tomaz MA, Lidon FC, DaMatta FM, Ramalho JC. Combined effects of elevated [CO2] and high temperature on leaf mineral balance in Coffea spp. plants. Climatic Change. 2014;126:365-379. DOI: 10.1007/s10584-014-1236-7
  35. 35. Rodrigues WP, Martins MQ, Fortunato AS, Rodrigues AP, Semedo JN, Simões-Costa MC, Pais IP, Leitão AE, Colwell F, Goulao L, Máguas C, Maia R, Partelli FL, Campostrini E, Scotti-Campos P, Ribeiro-Barros AI, Lidon FC, Damatta FM, Ramalho JC. Long-term elevated air [CO2] strengthens photosynthetic functioning and mitigates the impact of supra-optimal temperatures in tropical Coffea arabica and C. canephora species. Global Change Biology. 2016;22:415-431. DOI: 10.1111/gcb.13088
  36. 36. Verhage FYF, Antem NPR, Sentelhas PC. Carbon dioxide fertilization offsets negative impacts of climate change on Arabica coffee yield in Brazil. Climatic Change. 2017;144:671-685. DOI: 10.1007/s10584-017-2068-z
  37. 37. DaMatta FM., Rahn E, Laderach P, Ghini R, Ramalho JC. Why could the coffee crop endure climate change and global warming to a greater extent than previously estimated? Clim Change, 2018. (submitted)
  38. 38. Nunes MA. Water relations in coffee: Significance of plant water deficits to growth and yield: A review. Journal of Coffee Research. 1976;6:4-21. Available at: http://www.sidalc.net/cgi-bin/wxis.exe/?IsisScript=orton.xis&method=post&formato=2&cantidad=1&expresion=mfn=032673
  39. 39. Gutiérrez MV, Meinzer FC, Grantz DA. Regulation of transpiration in coffee hedgerows: Covariation of environmental variables and apparent responses of stomata to wind and humidity. Plant Cell Environment. 1994;17:1305-1313. DOI: 10.1111/j.1365-3040.1994.tb00532.x
  40. 40. Tausend PC, Goldstein G, Meinzer FC. Water utilization, plant hydraulic properties and xylem vulnerability in three contrasting coffee (Coffea arabica) cultivars. Tree Physiology. 2000;20:159-168. DOI: 10.1093/tree.phys/20.3.159
  41. 41. Lima ALS, DaMatta FM, Pinheiro HA, Totola MR, Loureiro ME. Photochemical responses and oxidative stress in two clones of Coffea canephora under water deficit conditions. Environmetal and Experimental Botany. 2002;47:239-247. DOI: https://doi.org/10.1016/S0098-8472(01)00130-7
  42. 42. DaMatta FM, Chaves ARM, Pinheiro HA, Ducatti C, Loureiro ME. Drought tolerance of two field-grown clones of Coffea canephora. Plant Science. 2003;164:111-117. DOI: https://doi.org/10.1016/S0168-9452(02)00342-4
  43. 43. Pinheiro HA, DaMatta FM, Chaves AR, Loureiro ME, Ducatti C. Drought tolerance is associated with rooting depth and stomatal control of water use in clones of Coffea canephora. Annals of Botany. 2005;96:101-108. DOI: https://doi.org/10.1093/aob/mci154
  44. 44. Nardini A, Ôunapuu-Pikas E, Savi T. When smaller is better: Leaf hydraulic conductance and drought vulnerability correlate to leaf size and venation density across four Coffea arabica genotypes. Functional Plant Biology. 2014;41:972-982. DOI: 10.1071/FP13302
  45. 45. Menezes-Silva PE, Cavatte PC, Martins SCV, Reis JV, Pereira LF, Ávila RT, Almeida AL, Ventrella MC, DaMatta FM. Wood density, but not leaf hydraulic architecture, is associated with drought tolerance in clones of Coffea canephora. Trees. 2015;29:1687-1697. DOI: 10.1007/s00468-015-1249-5
  46. 46. Miniussi M, Terra LD, Savi T, Pallavicini A, Nardini A. Aquaporins in Coffea arabica L.: Identification, expression, and impacts on plant water relations and hydraulics. Plant Physiology and Biochemistry. 2015;95:92-102. DOI: 10.1016/j.plaphy.2015.07.024
  47. 47. Moffato LS, Carneiro FA, Vieira NG, Duarte KE, Vidal RO, Alekcevetch JC, Cotta MG, Verdeil J-L, Lapeyre-Montes F, Lartad M, Leroy T, Bellis FD, Pot D, Rodrigues GC, Carazzolle MF, Pereira GAG, Andrade AC, Marraccini P. Identification of candidate genes for drought tolerance in coffee by high-throughput sequencing in the shoot apex of different Coffea arabica cultivars. BMC Plant Biology. 2016;16:94. DOI: https://doi.org/10.1186/s12870-016-0777-5
  48. 48. Menezes-Silva PE, Sanglard LMPV, Ávila RT, Morais LE, Martins SCV, Nobres P, Patreze CM, Ferreira MA, Araújo WL, Fernie AR, DaMatta FM. Photosynthetic and metabolic acclimation to repeated drought events play key roles in drought tolerance in coffee. Journal of Experimental Botany. 2017;68:4309-4322. DOI: 10.1093/jxb/erx211
  49. 49. Martins MQ, Rodrigues WP, DaMatta FM, Ramalho JDC, Partelli FL. O Cafeeiro no contexto de alterações climáticas. In: Ferreira a, Partelli FL, do Amaral JAT, Dalvi LP, Caldeira MVW, Coelho RI. Tópicos Especiais Em Genética e Melhoramento. Alegre: Suprema Gráfica e Editora; 2016. pp. 263-288
  50. 50. ICO (International Coffee Organization). World Coffee Trade (1963–2013): A Review of the Markets, Challenges and Opportunities Facing the Sector. ICC (International Coffee Council) 111-5 Rev. 1. 2014. 29 p. Available at http://www.ico.org/news/icc-111-5-r1e-world-coffee-outlook.pdf
  51. 51. Ramalho JC, DaMatta FM, Rodrigues AP, Scotti-Campos P, Pais I, Batista-Santos P, Partelli FL, Ribeiro A, Lidon FC, Leitão AE. Cold impact and acclimation response of Coffea spp. plants. Theoretical and Experimental Plant Physiology. 2014;26:5-18. DOI: 10.1007/s40626-014-0001-7
  52. 52. Waller JM, Bigger M, Hillocks RJ. World coffee production. In: Waller J, editor. Coffee Pests, Diseases and Their Management. Chapter 2. Egham, Surrey, UK: CAB International; 2007. pp. 17-40. DOI: 10.1079/9781845931292.0017
  53. 53. Osorio N. The Global Coffee Crisis: A Threat to Sustainable Development. London: International Coffee Organization; 2002. p. 4
  54. 54. ICO (International Coffee Organization). Trade statistics. Available from: http://www.ico.org/historical/1990%20onwards/PDF/1a-total-production.pdf [Accessed: 8 September 2017]
  55. 55. Bagyaraj DJ, Thilagar G, Ravisha C, Kushalappa CG, Krishnamurthy KN, Vaast P. Below ground microbial diversity as influenced by coffee agroforestry systems in the Western Ghats, India. Agriculture, Ecosystems and Environment. 2015;202:198-202. DOI: https://doi.org/10.1016/j.agee.2015.01.015
  56. 56. Krishnan S, Ranker TA, Davis AP, Rakotomalala JJ. An assessment of the genetic integrity of ex situ germplasm collections of three endangered species of Coffea from Madagascar: Implications for the management of field germplasm collections. Genetic Resources Crop Evolution. 2013;60:1021-1036. DOI: 10.1007/s10722-012-9898-3
  57. 57. Partelli FL, Batista-Santos P, Scotti-Campos P, Pais IP, Quartin VL, Vieira HD, Ramalho JC. Characterization of the main lipid components of chloroplast membranes and cold induced changes in Coffea spp. Environmental and Experimental Botany. 2011;74:194-204. DOI: 10.1016/j.envexpbot.2011.06.001
  58. 58. Davis AP, Tosh J, Ruch N, Fay MF. Growing coffee: Psilanthus (Rubiaceae) subsumed on the basis of molecular and morphological data implications for the size, morphology, distribution and evolutionary history of Coffea. Botanical Journal of the Linnean Society. 2011;167:357-377. DOI: 10.1111/j.1095-8339.2011.01177.x
  59. 59. Farah A, Monteiro MC, Calado V, Franca AS, Trugo LC. Correlation between cup quality and chemical attributes of Brazilian coffee. Food Chemistry. 2006;98:373-380. DOI: 10.1016/j.foodchem.2005.07.032
  60. 60. Joët T, Laffargue A, Descroix F, Doulbeau S, Bertrand B, Kochko A, Dussert S. Influence of environmental factors, wet processing and their interactions on the biochemical composition of green Arabica coffee beans. Food Chemistry. 2010;118:693-701. DOI: 10.1016/j.foodchem.2009.05.048
  61. 61. Bicho NC, Leitão AE, Ramalho JC, Alvarenga NBD, Lidon FC. Identification of nutritional descriptors of roasting intensity in beverages of Arabica and Robusta coffee beans. International Journal of Food Sciences and Nutrition. 2011;62:865-871. DOI: 10.3109/09637486.2011.588594
  62. 62. Bertrand B, Boulanger R, Dussert S, Ribeyre F, Berthiot L, Descroix F, Joët T. Climatic factors directly impact the volatile organic compound fingerprint in green Arabica coffee bean as well as coffee beverage quality. Food Chemistry. 2012;135:2575-2583. DOI: 10.1016/j.foodchem.2012.06.060
  63. 63. Ribeiro VS, Leitão AE, Ramalho JC, Lidon FC. Chemical characterization and antioxidant properties of a new coffee blend with cocoa, coffee silverskin and green coffee minimally processed. Food Research International. 2014;61:39-47. DOI: http://dx.doi.org/10.1016/j.foodres.2014.05.003
  64. 64. Cheng B, Furtado A, Smyth HE, Henry RJ. Influence of genotype and environment on coffee quality. Trends in Food Science & Technology. 2016;57:20-30. DOI: http://dx.doi.org/10.1016/j.tifs.2016.09.003
  65. 65. Camargo AP. O clima e a cafeicultura no Brasil. Informe Agropecuário. 1985;11:13-26. ISBN: 01003364
  66. 66. Scotti-Campos P, Pais IP, Partelli FL, Batista-Santos P, Ramalho JC. Phospholipids profile in chloroplasts of Coffea spp. genotypes differing in cold acclimation ability. Journal of Plant Physiology. 2014;171:243-249. DOI: https://doi.org/10.1016/j.jplph.2013.07.007
  67. 67. Charrier A, Berthaud J. Botanical classification of coffee. In: Clifforf MN, Willson KC, editors. Coffee: Botany, Biochemistry, and Production of Beans and Beverage. London: Crom Helm; 1985. pp. 13-47
  68. 68. Coste R. Coffee: The Plant and the Product. London: MacMillan Press Ltd; 1992. 328 p
  69. 69. Davis AP, Govaerts R, Bridson DM, Stoffelen P. An annotated taxonomic conspectus of the genus Coffea (Rubiaceae). Botanical Journal of the Linnean Society. 2006;152:465-512. DOI: 10.1111/j.1095-8339.2006.00584.x
  70. 70. Partelli FL, Vieira HD, Silva MG, Ramalho JC. Seasonal vegetative growth of different age branches of conilon coffee tree. Semina: Ciências Agrárias. 2010;31:619-626. DOI: 10.5433/1679-0359.2010v31n3p619
  71. 71. Partelli FL, Marré WB, Falqueto AR, Vieira HD, Cavatti PC. Seasonal vegetative growth in genotypes of Coffea canephora, as related to climatic factors. Journal of Agricultural Science. 2013;5:108-116. DOI: http://dx.doi.org/10.5539/jas.v5n8p108
  72. 72. Covre AM, Partelli FL, Bonomo R, Braun H, Ronchi CP. Vegetative growth of Conilon coffee plants under two water conditions in the Atlantic region of Bahia State, Brazil. Acta Scientiarum. Agronomy. 2016;38:535-545. DOI: http://dx.doi.org/10.4025/actasciagron.v38i4.30627
  73. 73. Dubberstein D, Partelli FL, Dias JRM, Espindula MC. Influence of fertilization on coffee trees vegetative growth in west south Amazon. Coffee Science. 2017;12:197-206. Available at: http://www.coffeescience.ufla.br/index.php/Coffeescience/article/view/1228
  74. 74. Taques RC, Dadalto GG. Zoneamento agroclimático para a cultura do café Conilon no Estado do Espírito Santo. In: Ferrão RG, Fonseca AFA, Ferrão MAG, De Muner LH. Café Conilon. 2nd ed. Vitória: Incaper; 2017. pp. 69-79
  75. 75. Xu Z, Jiang Y, Zhou G. Response and adaptation of photosynthesis, respiration, and antioxidant systems to elevated CO2 with environmental stress in plants. Frontiers in Plant Science. 2015;6:701-717. DOI: https://doi.org/10.3389/fpls.2015.00701
  76. 76. Jiang Y, Xu Z, Zhou G, Liu T. Elevated CO2 can modify the response to a water status gradient in a steppe grass: From cell organelles to photosynthetic capacity to plant growth. BMC Plant Biology. 2016;16:157. DOI: 10.1186/s12870-016-0846-9
  77. 77. Song Y, Yu J, Huang B. Elevated CO2-mitigation of high temperature stress associated with maintenance of positive carbon balance and carbohydrate accumulation in Kentucky bluegrass. PLoS One. 2014;9:e89725. DOI: https://doi.org/10.1371/journal.pone.0089725
  78. 78. Djanaguiraman M, Vara Prasad PV. Ethylene production under high temperature stress causes premature leaf senescence in soybean. Functional Plant Biology. 2010;37:1071-1084. DOI: 10.1071/FP10089
  79. 79. Wise RR, Olson AJ, Schrader SM, Sharkey TD. Electron transport is the functional limitation of photosynthesis in field-grown pima cotton plants at high temperature. Plant Cell & Environment. 2004;27:717-724. DOI: 10.1111/j.1365-3040.2004.01171.x
  80. 80. Wahid A, Gelani S, Ashraf M, Foolad MR. Heat tolerance in plants: An overview. Environmental and Experimental Botany. 2007;61:199-223. DOI: 10.1016/j.envexpbot.2007.05.011
  81. 81. Lambers H, Chapin III FS, Pons TL. Plant Physiological Ecology. 2nd ed. New York: Springer-Verlag; 2008. 604 p. DOI: 10.1007/978-0-387-78341-3
  82. 82. Crafts-Brandner SJ, Salvucci ME. Rubisco activase constrains the photosynthetic potential of leaves at high temperature and CO2. Proceedings of the National Academy of Sciences of the United States of America. 2000;97:13430-13435. DOI: 10.1073/pnas.230451497
  83. 83. Ainsworth EA, Rogers A. The response of photosynthesis and stomatal conductance to rising [CO2]: Mechanisms and environmental interactions. Plant Cell & Environment. 2007;30:258-270. DOI: 10.1111/j.1365-3040.2007.01641.x
  84. 84. Barnabás B, Jäger K, Fehér A. The effect of drought and heat stress on reproductive processes in cereals. Plant Cell Environment. 2008;31:11-38. DOI: 10.1111/j.1365-3040.2007.01727.x
  85. 85. Assad ED, Pinto HS, Zullo Jr J, Ávila AMH. Impacto das mudanças climáticas no zoneamento agroclimático do café no Brasil. Pesquisa Agropecuária Brasileira. 2004;39:1057-1064. DOI: http://dx.doi.org/10.1590/S0100-204X2004001100001
  86. 86. Martins MQ, Fortunato AS, Rodrigues WP, Partelli FL, Campostrini E, Lidon FC, DaMatta FM, Ramalho JC, Ribeiro-Barros AI. Selection and validation of reference genes for accurate RT-qPCR data normalization in Coffea spp. under a climate changes context of interacting elevated [CO2] and temperature. Frontier. Plant Science. 2017;8:307. DOI: https://doi.org/10.3389/fpls.2017.00307
  87. 87. Barros R, Mota JWS, DaMatta FM, Maestri M. Decline of vegetative growth in Coffea arabica L. in relation to leaf temperature: Water potential and stomatal conductance. Field Crops Research. 1997;54:65-72. DOI: https://doi.org/10.1016/S0378-4290(97)00045-2
  88. 88. Rodrigues WP, Machado Filho JA, Silva JR, Figueiredo FAMMA, Ferraz TM, Ferreira LS, Bezerra LS, Abreu DP, Bernado WP, Cespom L, Sousa EF, Glenn DM, Ramalho JC, Campostrini E. Whole-canopy gas exchange in Coffea sp. is affected by supra-optimal temperature and light distribution within the canopy: The insights from an improved multi-chamber system. Scientia Horticulturae. 2016;211:194-202. DOI: https://doi.org/10.1016/j.scienta.2016.08.022
  89. 89. Thioune EH, McCarthy J, Gallagher T, Osborne BA. Humidity shock leads to rapid, temperature dependent changes in coffee leaf physiology and gene expression. Tree Physiology. 2017;37:367-379. DOI: 10.1093/treephys/tpw129
  90. 90. Noctor G, Mhamdi A, Foyer CH. The roles of reactive oxygen metabolism in drought: Not so cut and dried. Plant Physiology. 2014;164:1636-1648. DOI: 10.1104/pp.113.233478
  91. 91. Barbosa MR, Silva MMDA, Willadino L, Ulisses C, Camara TR. Geração e desintoxicação enzimática de espécies reativas de oxigênio em plantas. Ciência Rural. 2014;44:453-460. DOI: http://dx.doi.org/10.1590/S0103-84782014000300011
  92. 92. Pinheiro HA, DaMatta FM, Chaves ARM, Loureiro ME, Ducatti C. Drought tolerance in relation to protection against oxidative stress in clones of Coffea canephora subjected to long-term drought. Plant Science. 2004;167:1307-1314. DOI: https://doi.org/10.1016/j.plantsci.2004.06.027
  93. 93. Martins SCV, Araújo WA, Tohge T, Fernie AR, DaMatta FM. High-light-acclimated coffee plants the metabolic machinery is adjusted to avoid oxidative stress rather than to benefit from extra light enhancement in photosynthetic yield. PLoS One. 2014;9:e94862. DOI: https://doi.org/10.1371/journal.pone.0094862
  94. 94. Trueba S, Isnard S, Barthélémy D, Olson ME. Trait coordination, mechanical behaviour and growth form plasticity of Amborella trichopoda under variation in canopy openness. AoB Plants. 2016;8:plw068. DOI: 10.1093/aobpla/plw068
  95. 95. Meinzer FC, Goldstein G, Grantz DA. Carbon isotope discrimination in coffee genotypes grown under limited water supply. Plant Physiology. 1990;92:130-135. DOI: 0032-0889/90/92/0130/06/$01.00/0
  96. 96. Zhu C, Ziska L, Zhu J, Zeng Q, Xie Z, Tang H, Jia X, Hasegawa T. The temporal and species dynamics of photosynthetic acclimation in flag leaves of rice (Oryza sativa) and wheat (Triticum aestivum) under elevated carbon dioxide. Physiologia Plantarum. 2012;145:395-405. DOI: 10.1111/j.1399-3054.2012.01581.x
  97. 97. Long SP, Ainsworth EA, Rogers A, Ort DR. Rising atmospheric carbon dioxide: Plants FACE the future. Annual Review Plant Biology. 2004;55:591-628. DOI: 10.1146/annurev.arplant.55.031903.141610
  98. 98. Rakocevic M, Ferrandes R, Marchiori PER, Ribeiro RV. Estimating the canopy architecture and photosynthesis of Coffea arabica L. plants cultivated under long-term elevated air CO2 concentration. In: International Conference on Functional-Structural Plant Growth Modeling, Simulation, Visualization and Applications (FSPMA); 7–11 Nov 2016; Qingdao. New York: IEEE; 2017. pp. 175-182. DOI: 10.1109/FSPMA.2016.7818304
  99. 99. Wertin TM, McGuire MA, Teskey RO. The influence of elevated temperature, elevated atmospheric CO2 concentration and water stress on net photosynthesis of loblolly pine (Pinus taeda L.) at northern, central and southern sites in its native range. Global Change Biology. 2010;16:2089-2103. DOI: 10.1111/j.1365-2486.2009.02053.x
  100. 100. Nunes MA, Ramalho JC, Dias MA. Effect of nitrogen supply on the photosynthetic performance of leaves from coffee plants exposed to bright light. Journal of Experimental Botany. 1993;44:893-899. DOI: 10.1093/jxb/44.5.893
  101. 101. Ramalho JC, Campos PS, Teixeira M, Nunes MA. Nitrogen dependent changes in antioxidant systems and in fatty acid composition of chloroplast membranes from Coffea arabica L. plants submitted to high irradiance. Plant Science. 1998;135:115-124. DOI: https://doi.org/10.1016/S0168-9452(98)00073-9
  102. 102. Ramalho JC, Pons T, Groeneveld H, Azinheira HG, Nunes MA. Photosynthetic acclimation to high light conditions in mature leaves of Coffea arabica L.: Role of xanthophylls, quenching mechanisms and nitrogen nutrition. Functional Plant Biology. 2000;27:43-51. DOI: 10.1071/PP99013
  103. 103. Ramalho JC, Fortunato AS, Goulao LF, Lidon FC. Cold-induced changes in mineral content in Coffea spp. leaves – identification of descriptors for tolerance assessment. Biologia Plantarum. 2013;57:495-506. DOI: 10.1007/s10535-013-0329-x
  104. 104. Ramalho JC, Rebelo MC, Santos ME, Antunes ML, Nunes MA. Effects of calcium deficiency on Coffea arabica. Nutrient changes and correlation of calcium levels with some photosynthetic parameters. Plant and Soil. 1995;172:87-96. DOI: 10.1007/BF00020862
  105. 105. Blank RR, Morgan T, Ziska LH, White RH. Effect of atmospheric CO2 levels on nutrients in cheatgrass tissue. Natural Resources and Environmental Issues. 2011;16:1-6. Available form: http://digitalcommons.usu.edu/nrei/vol16/iss1/18
  106. 106. Chaturvedi AK, Bahuguna RN, Pal M, Shah D, Maurya S, Jagadish KSV. Elevated CO2 and heat stress interactions affect grain yield, qualityand mineral nutrient composition in rice under field conditions. Field Crops Research. 2017;206:149-157. DOI: http://dx.doi.org/10.1016/j.fcr.2017.02.018
  107. 107. Conroy J, Hocking P. Nitrogen nutrition of C3 plants at elevated atmospheric CO2 concentrations. Physiologia Plantarum. 1993;89:570-576. DOI: 10.1111/j.1399-3054.1993.tb05215.x
  108. 108. Thiec DL, Dixon M, Loosveldt P, Garrec JP. Seasonal and annual variations of phosphorus, calcium, potassium and manganese contents in different cross-sections of Picea abies (L.) karst. Needles and Quercus rubra L. leaves exposed to elevated CO2. Trees. 1995;10:55-62. DOI: https://doi.org/10.1007/BF00192184
  109. 109. Cotrufo MF, Ineson P, Scott A. Elevated CO2 reduces the nitrogen concentration of plant tissues. Global Change Biology. 1998;4:43-54. DOI: 10.1046/j.1365-2486.1998.00101.x
  110. 110. Overdieck D. Elevated CO2 and the mineral content of herbaceous and woody plants. Vegetatio. 1993;104:403-411. DOI: https://doi.org/10.1007/BF00048169
  111. 111. Manderscheid R, Bender J, Jäger H-J, Weigel HJ. Effects of season long CO2 enrichment on cereals. II. Nutrient concentrations and grain quality. Agriculture, Ecosystems & Environment. 1995;54:175-185. DOI: https://doi.org/10.1016/0167-8809(95)00602-O
  112. 112. Fangmeier A, Grüters U, Högy P, Vermehren B, Jäger HJ. Effects of elevated CO2, nitrogen supply, and tropospheric ozone on spring wheat. II. Nutrients (N, P, K, S, Ca, mg, Fe, Mn, Zn). Environmental Pollution. 1997;96:43-59. DOI: 10.1016/S0269-7491(97)00013-4
  113. 113. Leakey ADB, Ainsworth EA, Bernacchi CJ, Rogers A, Long SP, Ort DR. Elevated CO2 effects on plant carbon, nitrogen, and water relations: Six important lessons from FACE. Journal of Experimental Botany. 2009;60:2859-2876. DOI: https://doi.org/10.1093/jxb/erp096
  114. 114. Fortunato A, Lidon FC, Batista-Santos P, Leitão AE, Pais IP, Ribeiro AI, Ramalho JC. Biochemical and molecular characterization of the antioxidative system of Coffea sp. under cold conditions in genotypes with contrasting tolerance. Journal of Plant Physiology. 2010;167:333-342. DOI: 10.1016/j.jplph.2009.10.013
  115. 115. DaMatta FM, Ronchi CP, Maestri M, Barros RS. Ecophysiology of coffee growth and production. Brazilian Journal of Plant Physiology. 2008;19:485-510. DOI: http://dx.doi.org/10.1590/S1677-04202007000400014
  116. 116. Kumar D, Tieszen LL. Photosynthesis in Coffea arabica. L. Effects of light and temperature. Experimental Agriculture. 1980;16:13-19. DOI: 10.1017/S0014479700010656
  117. 117. Rodríguez-López CPC, Menezes-Silva PE, Martins SCV, Morais LE, Medina EF, DaMatta FM. Physiological and biochemical abilities of robusta coffee leaves for acclimation to cope with temporal changesin light availability. Physiologia Plantarum. 2013;149:45-55. DOI: 10.1111/ppl.12010
  118. 118. Cavatte PC, Oliveira AA, Morais LE, Martins SC, Sanglard LM, DaMatta FM. Could shading reduce the negative impacts of drought on coffee? A morphophysiological analysis. Physiologia Plantarum. 2012;114:111-122. DOI: 10.1111/j.1399-3054.2011.01525.x
  119. 119. Beer J. Advantages, disadvantages and desirable characteristics of shade trees for coffee, cocoa and tea. Agroforesty Systemys. 1987;5:3-13. DOI: https://doi.org/10.1007/BF00046410
  120. 120. Souza HN, Graaff J, Pullemam MM. Strategies and economics of farming systems with coffee in the Atlantic rainforest biome. Agroforestry Systems. 2012;84:227-242. DOI: https://doi.org/10.1007/s10457-011-9452-x
  121. 121. Barradas VL, Fanjul L. Microclimatic chacterization of shaded and open-grown coffee (Coffea arabica L.) plantations in Mexico. Agricultural and Foresty Metereology. 1986;38:101-112. DOI: https://doi.org/10.1016/0168-1923(86)90052-3
  122. 122. Vaast P, Angrand J, Franck N, Dauzat J, Genard M. Fruit load and branch ring-barking affect carbon allocation and photosynthesis of leaf and fruit of Coffea arabica in the field. Tree Physiology. 2005;25:753-760. DOI: 10.1093/treephys/25.6.753
  123. 123. Siebert SF. From shade- to sun-grown perennial crops in Sulawesi, Indonesia: Implications for biodiversity conservation and soil fertility. Biodiversity and Conservation. 2002;11:1889-1902. DOI: https://doi.org/10.1023/A:1020804611740
  124. 124. Scherr SJ, McNeely JA. Biodiversity conservation and agricultural sustainability: Towards a new paradigm of ‘ecoagriculture’ landscapes. Philosophical Transactions of the Royal Society. 2008;363:477-494. DOI: 10.1098/rstb.2007.2165
  125. 125. Siles P, Harmand JM, Vaast P. Effects of Inga densiflora on the microclimate of coffee (Coffea arabica L.) and overall biomass under optimal growing conditions in Costa Rica. Agroforestry Systems. 2010;78:269-286. DOI: 10.1007/s10457-009-9241-y
  126. 126. Charbonnier F, Roupsard O, Maire GL, Guillemot J, Casanoves F, Lacointe A, Vaast P, Allinne C, Audebert L, Cambou A, Clément-Vidal A, Defrenet E, Duursma RA, Jarri L, Jourdan C, Khac E, Leandro P, Medlyn BE, Saint-André L, Thaler P, Meersche KVD, Aguilar AB, Lehner P, Dreyer E. Increased light-use efficiency sustains net primary productivity of shaded coffee plants in agroforestry system. Plant, Cell and Environment. 2017;40:1592-1608. DOI: 10.1111/pce.12964
  127. 127. Bragança SM, Carvalho CHS, Fonseca AFA, Ferrão RG. Clonal varieties of Conilon coffee for the Espírito Santo state, Brazil. Pesquisa Agropecuária Brasileira. 2001;36:765-770. DOI: http://dx.doi.org/10.1590/S0100-204X2001000500006
  128. 128. Oliosi G, Giles JD, Rodrigues WP, Ramalho JC, Partelli LP. Microclimate and development of Coffea canephora cv. Conilon under different shading levels promoted by Australian cedar (Toona ciliate M. Roem. Var. Australis). Australian Journal of Crop Science. 2016;10:528-538. DOI: 10.21475/ajcs.2016.10.04.p7295x
  129. 129. Partelli FL, Araújo AV, Vieira HD, Dias JRM, Menezes LFT, Ramalho JC. Microclimate and development of ‘Conilon’ coffee intercropped with rubber trees. Pesquisa Agropecuária Brasileira. 2014;49:872-881. DOI: http://dx.doi.org/10.1590/S0100-204X2014001100006
  130. 130. Ricci MSF, Costa JR, Pinto AN, Santos VLS. Organic cultivation of coffee cultivars grown under full sun and under shading. Pesquisa Agropecuária Brasileira. 2006;41:569-575. DOI: http://dx.doi.org/10.1590/S0100-204X2006000400004
  131. 131. Somporn C, Kamtuo A, Theerakulpisut P, Siriamornpun S. Effect of shading on yield, sugar content, phenolic acids and antioxidant property of coffee beans (Coffea arabica L. cv. Catimor) harvested from North-Eastern Thailand. Journal of the Science of Food and Agriculture. 2012;92:1956-1963. DOI: 10.1002/jsfa.5568
  132. 132. Muschler RG. Shade improves coffee quality in a sub-optimal coffee zone of Costa Rica. Agroforesty Systems. 2001;85:131-139. DOI: http://dx.doi.org/10.1023/A:1011863426305
  133. 133. Vaast P, Bertrand B, Perriot JJ, Guyot B, Genard M. Fruit thinning and shade improve bean characteristics and beverage quality of coffee (Coffea arabica L.) under optimal conditions. Journal of the Science of Food and Agriculture. 2006;86:197-204. DOI: 10.1002/jsfa.2338
  134. 134. Camargo APD, Camargo MBD. Definition and outline for the phenological phases of arabica coffee under Brazilian tropical conditions. Bragantia. 2001;60:65-68. DOI: http://dx.doi.org/10.1590/S0006-87052001000100008
  135. 135. Glenn DM, Prado E, Erez A, McFerson J, Puterka GJA. Reflective, processed kaolin particle film affects fruit temperature, radiation reflection, and solar injury in apple. Journal of the American Society for Horticultural Science. 2002;127:188-193. Available from: https://www.ars.usda.gov/ARSUserFiles/2017/Sunburn%20paper%20ASHS.pdf
  136. 136. Glenn DM. Particle film mechanisms of action that reduce the effect of environmental stress in “Empire” apple. Journal of the American Society for Horticultural Science. 2009;134:314-321. Available at: http://journal.ashspublications.org/content/134/3/314.full
  137. 137. Glenn DM, Cooley N, Shellie K. Impact of kaolin particle film and water deficit on wine grape water use efficiency and plant water relations. Hortscience. 2010;45:1178-1187. Available at: http://hortsci.ashspublications.org/content/45/8/1178.abstract
  138. 138. Brilhante L, Belfiore N, Gaiotti F, Lovat L, Sansone L, Poni S, Tomasi D. Comparing kaolin and pinolene to improve sustainable grapevine production during drought. PLoS One. 2016;110:e0156631. DOI: 10.1371/journal.pone.0156631
  139. 139. Steiman SR, Bittenbender HB. Analysis of kaolin particle film use and its application on coffee. Hortscience. 2007;42:1605-1608. Available from: http://hortsci.ashspublications.org/content/42/7/1605.full
  140. 140. Verdin-Filho AC, Tomaz MA, Ferrão RG, Ferrão MAG, Fonseca AFA, Rodrigues WN. Conilon coffee yield using the programmed pruning cycle different cultivation densities. Coffee Science. 2014;9:489-494. Available at: http://www.coffeescience.ufla.br/index.php/Coffeescience/article/download/734/pdf_130
  141. 141. Zuazo VHD, Pleguezuelo CRR. Soil-erosion and runoff prevention by plant covers. A review. Agronomy for Sustainable Development. 2008;28:65-86. DOI: 10.1051/agro:2007062
  142. 142. Santana AO, Cuniat G, Imaña-Encinas J. Contribution of understory vegetation on total evapotranspiration in distinct environments of the cerrado biome, Distrito Federal, Brazil. Ciência Florestal. 2010;20:269-281. DOI: http://dx.doi.org/10.5902/198050981851
  143. 143. Alcântara EN, Ferreira MM. Effects of weed control methods in coffee (Coffea arabica L.) on soil physical quality. Revista Brasileira de Ciência do Solo. 2000;24:711-721. DOI: http://dx.doi.org/10.1590/S0100-06832000000400003
  144. 144. Alcântara EN de, Nobrega JCA, Ferreira MM. Métodos de controle de plantas daninhas no cafeeiro afetam os atributos químicos do solo. Ciência Rural. 2009;39:749-757. DOI: http://dx.doi.org/10.1590/S0103-84782009000300018
  145. 145. Effgen TAM, Passos RR, Andrade FV, Lima JSS, Reis EF, Borges EN. Physical soil properties as a function of management in crops of conilon coffee. Revista Ceres. 2012;59:414-421. DOI: 10.1590/S0034-737X2012000300018
  146. 146. Fernandes ALT, Santinato F, Ferreira RT, Santinado R. Reduction of mineral fertilization of arabic coffee using coffee straw. Coffee Science. 2013;8:324-336. Available from: http://www.coffeescience.ufla.br/index.php/Coffeescience/article/view/454/pdf_44
  147. 147. Paes JMV, Andreola F, Brito CH, Loures EG, Decomposição d. Palha de café em três tipos de solo e sua influência sobre a CTC e o pH. Revista Ceres. 1996;43:674-683. Available from: http://www.ceres.ufv.br/ojs/index.php/ceres/article/viewFile/2373/381
  148. 148. Almeida FS. Efeitos alelopáticos de resíduos vegetais. Pesquisa Agropecuária Brasileira. 1991;26:221-236. Available from: https://seer.sct.embrapa.br/index.php/pab/article/view/3333/666
  149. 149. Sharma E, Rai SC, Sharma R. Soil, water and nutrient conservation in mountain farming systems: Case-study from the Sikkim Himalaya. Journal of Environmental Management. 2001;61:123-135. DOI: 10.1006/jema.2000.0386
  150. 150. Hammad AA, Haugen LE, Borresen T. Effects of stonewalled terracing techniques on soil-water conservation and wheat production under mediterranean conditions. Environmental Management. 2004;34:701-710. DOI: 10.1007/s00267-003-0278-9
  151. 151. Al-Seekh SH, Mohammad AG. The effect of water harvesting techniques on runoff, sedimentation, and soil properties. Environmental Management. 2009;44:37-45. DOI: 10.1007/s00267-009-9310-z
  152. 152. Magalhães GMF. Analysis of efficiency of retention terraces in sub-basins of the São Francisco River. Revista Brasileira de Engenharia Agrícola e Ambiental. 2013;17:1109-1115. DOI: http://dx.doi.org/10.1590/S1415-43662013001000013

Written By

Danielly Dubberstein, Weverton P. Rodrigues, José N. Semedo, Ana P. Rodrigues, Isabel P. Pais, António E. Leitão, Fábio L. Partelli, Eliemar Campostrini, Fernando Reboredo, Paula Scotti-Campos, Fernando C. Lidon, Ana I. Ribeiro-Barros, Fábio M. DaMatta and José C. Ramalho

Submitted: 27 June 2017 Reviewed: 10 November 2017 Published: 20 December 2017