U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Front Plant Sci

Climate Change and the Impact of Greenhouse Gasses: CO 2 and NO, Friends and Foes of Plant Oxidative Stress

Here, we review information on how plants face redox imbalance caused by climate change, and focus on the role of nitric oxide (NO) in this response. Life on Earth is possible thanks to greenhouse effect. Without it, temperature on Earth’s surface would be around -19°C, instead of the current average of 14°C. Greenhouse effect is produced by greenhouse gasses (GHG) like water vapor, carbon dioxide (CO 2 ), methane (CH 4 ), nitrous oxides (N x O) and ozone (O 3 ). GHG have natural and anthropogenic origin. However, increasing GHG provokes extreme climate changes such as floods, droughts and heat, which induce reactive oxygen species (ROS) and oxidative stress in plants. The main sources of ROS in stress conditions are: augmented photorespiration, NADPH oxidase (NOX) activity, β-oxidation of fatty acids and disorders in the electron transport chains of mitochondria and chloroplasts. Plants have developed an antioxidant machinery that includes the activity of ROS detoxifying enzymes [e.g., superoxide dismutase (SOD), ascorbate peroxidase (APX), catalase (CAT), glutathione peroxidase (GPX), and peroxiredoxin (PRX)], as well as antioxidant molecules such as ascorbic acid (ASC) and glutathione (GSH) that are present in almost all subcellular compartments. CO 2 and NO help to maintain the redox equilibrium. Higher CO 2 concentrations increase the photosynthesis through the CO 2 -unsaturated Rubisco activity. But Rubisco photorespiration and NOX activities could also augment ROS production. NO regulate the ROS concentration preserving balance among ROS, GSH, GSNO, and ASC. When ROS are in huge concentration, NO induces transcription and activity of SOD, APX, and CAT. However, when ROS are necessary (e.g., for pathogen resistance), NO may inhibit APX, CAT, and NOX activity by the S-nitrosylation of cysteine residues, favoring cell death. NO also regulates GSH concentration in several ways. NO may react with GSH to form GSNO, the NO cell reservoir and main source of S-nitrosylation. GSNO could be decomposed by the GSNO reductase (GSNOR) to GSSG which, in turn, is reduced to GSH by glutathione reductase (GR). GSNOR may be also inhibited by S-nitrosylation and GR activated by NO. In conclusion, NO plays a central role in the tolerance of plants to climate change.

Introduction

Life on Earth, as it is, relies on the natural atmospheric greenhouse effect. This is the result of a process in which a planet’s atmosphere traps the sun radiation and warms the planet’s surface.

Greenhouse effect occurs in the troposphere (the lower atmosphere layer), where life and weather occur. In the absence of greenhouse effect, the average temperature on Earth’s surface is estimated around -19°C, instead of the current average of 14°C ( Le Treut et al., 2007 ). Greenhouse effect is produced by greenhouse gasses (GHG). GHG are those gaseous constituents of the atmosphere that absorb and emit radiation in the thermal infrared range ( IPCC, 2014 ). Traces of GHG, both natural and anthropogenic, are present in the troposphere. The most abundant GHG in increasing order of importance are: water vapor, carbon dioxide (CO 2 ), methane (CH 4 ), nitrous oxides (N x O) and ozone (O 3 ) ( Kiehl and Trenberth, 1997 ). GHG percentages vary daily, seasonally, and annually.

GHG Contribute Differentially to Greenhouse Effect

Water vapor.

Water is present in the troposphere both as vapor and clouds. Water vapor was reported by Tyndal in 1861 as the most important gaseous absorber of variations in infrared radiation (cited in Held and Souden, 2000 ). Further accurate calculation estimate that water vapor and clouds are responsible for 49 and 25%, respectively, of the long wave (thermal) absorption ( Schmidt et al., 2010 ). However, atmospheric lifetime of water vapor is short (days) compared to other GHG as CO 2 (years) ( IPCC, 2014 ).

Water vapor concentrations are not directly influenced by anthropogenic activity and vary regionally. However, human activity increases global temperatures and water vapor formation indirectly, amplifying the warming in a process known as water vapor feedback ( Soden et al., 2005 ).

Carbon Dioxide (CO 2 )

Carbon dioxide is responsible for 20% of the thermal absorption ( Schmidt et al., 2010 ).

Natural sources of CO 2 include organic decomposition, ocean release and respiration. Anthropogenic CO 2 sources are derived from activities such as cement manufacturing, deforestation, fossil fuels combustion such as coal, oil and natural gas, etc. Surprisingly, 24% of direct CO 2 emission comes from agriculture, forestry and other land use, and 21% comes from industry ( IPCC, 2014 ).

Atmospheric CO 2 concentrations climbed up dramatically in the past two centuries, rising from around 270 μmol.mol -1 in 1750 to present concentrations higher than 385 μmol.mol -1 ( Mittler and Blumwald, 2010 ; IPCC, 2014 ). Around 50% of cumulative anthropogenic CO 2 emissions between 1750 and 2010 have taken place since the 1970s ( IPCC, 2014 ). It is calculated that the temperature rise produced by high CO 2 concentrations, plus the water positive feedback, would increase by 3–5°C the global mean surface temperature in 2100 ( IPCC, 2014 ).

Methane (CH 4 )

Methane (CH 4 ) is the main atmospheric organic trace gas. CH 4 is the primary component of natural gas, a worldwide fuel source. Significant emissions of CH 4 result from cattle farming and agriculture, but mainly as a consequence of fossil fuel use. Concentrations of CH 4 were multiplied by two since the pre-industrial era. The present worldwide-averaged concentration is of 1.8 μmol.mol -1 ( IPCC, 2014 ).

Although its concentration represents only 0.5% that of CO 2 , concerns arise regarding a jump in CH 4 atmospheric release. Indeed, it is 30 times more powerful than CO 2 as GHG ( IPCC, 2014 ). CH 4 generates O 3 (see below), and along with carbon monoxide (CO), contributes to control the amount of OH in the troposphere ( Wuebbles and Hayhoe, 2002 ).

Nitrous Oxides (NxO)

Nitrous oxide (N 2 O) and nitric oxide (NO) are GHG. During the last century, their global emissions have rised, due mainly to human intervention ( IPCC, 2014 ). The soil emits both N 2 O and NO. N 2 O is a strong GHG, whereas NO contributes indirectly to O 3 synthesis. As GHG, N 2 O is potentially 300 times stronger than CO 2 . Once in the stratosphere, the former catalyzes the elimination of O 3 ( IPCC, 2014 ). In the atmosphere, N 2 O concentrations are climbing up due mainly to microbial activity in nitrogen (N)-rich soils related with agricultural and fertilization practices ( Hall et al., 2008 ).

Anthropogenic emissions (from combustion of fossil fuels) and biogenic emissions from soils are the main sources of NO in the atmosphere ( Medinets et al., 2015 ). In the troposphere, NO quickly oxidizes to nitrogen dioxide (NO 2 ). NO and NO 2 (termed as NO x ) may react with volatile organic compounds (VOCs) and hydroxyl, resulting in organic nitrates and nitric acid, respectively. They access ecosystems through atmospheric deposition that has an impact on the N cycle as a result of acidification or N enrichment ( Pilegaard, 2013 ).

NO Sources and Chemical Reactions in Plants

Two major pathways for NO production have been described in plants: the reductive and the oxidative pathways. The reductive pathway involves the reduction of nitrite to NO by NR under conditions such as acidic pH, anoxia, or an increase in nitrite levels ( Rockel et al., 2002 ; Meyer et al., 2005 ). NR-dependent NO formation has been involved in processes such as stomatal closure, root development, germination and immune responses. In plants, nitrite may also be reduced enzymatically by other molybdenum enzymes such as, xanthine oxidase, aldehyde oxidase, and sulfite oxidase, in animals ( Chamizo-Ampudia et al., 2016 ) or via the electron transport system in mitochondria ( Gupta and Igamberdiev, 2016 ).

The oxidative pathway produces NO through the oxidation of organic compounds such as polyamines, hydroxylamine and arginine. In animals, NOS catalyzes arginine oxidation to citrulline and NO. Many efforts were made to find the arginine-dependent NO formation in plants, as well as of plant NOS ( Frohlich and Durner, 2011 ). The identification of NOS in the green alga Ostreococcus tauri ( Foresi et al., 2010 ) led to high-throughput bioinformatic analysis in plant genomes. This study shows that NOS homologs were not present in over 1,000 genomes of higher plants analyzed, but only in few photosynthetic microorganisms, such as algae and diatoms ( Di Dato et al., 2015 ; Kumar et al., 2015 ; Jeandroz et al., 2016 ). In summary, although an arginine-dependent NO production is found in higher plants, the specific enzyme/s involved in the oxidative pathways remain elusive.

Ozone (O 3 )

Ozone (O 3 ) is mainly found in the stratosphere, but a little amount is generated in the troposphere. Stratospheric ozone (namely the ozone layer) is formed naturally by chemical reactions involving solar ultraviolet (UV) radiation and O 2 . Solar UV radiation breaks one O 2 molecule, producing two oxygen atoms (2 O). Then, each of these highly reactive atoms combines with O 2 to produce an (O 3 ) molecule. Almost 99% of the Sun’s medium-frequency UV light (from about 200 to 315 nm wavelength) is absorbed by the (O 3 ) layer. Otherwise, they could damage exposed life forms near the Earth surface 1 .

The majority of tropospheric O 3 appears when NOx, CO and VOCs, react in the presence of sunlight. However, it was reported that NOx may scavenge O 3 in urban areas ( Gregg et al., 2003 ). This dual interaction between NOx and O 3 is influenced by light, season, temperature and VOC concentration ( Jhun et al., 2015 ).

Besides, the oxidation of CH 4 by OH in the troposphere gives way to formaldehyde (CH 2 O), CO, and O 3 , in the presence of high amounts of NOx 1 .

Tropospheric O 3 is harmful to both plants and animals (including humans). O 3 affects plants in several ways. Stomata are the cells, mostly on the underside of the plant leaves, that allow CO 2 and water to diffuse into the tissue. High concentrations of O 3 cause plants to close their stomata ( McAdam et al., 2017 ), slowing down photosynthesis and plant growth. O 3 may also provoke strong oxidative stress, damaging plant cells ( Vainonen and Kangasjärvi, 2015 ).

Global Climate Change: an Integrative Balance of the Impact on Plants

Anthropogenic activity alters global climate by interfering with the flows of energy through changes in atmospheric gasses composition, more than the actual generation of heat due to energy usage ( Karl and Trenberth, 2003 ). Short-term consequences of GHG increase in plants are mainly associated with the rise in atmospheric CO 2 . Plants respond directly to elevated CO 2 increasing net photosynthesis, and decreasing stomatal opening ( Long et al., 2004 ). To a lesser extent, O 3 uptake by plants may reduce photosynthesis and induce oxidative stress. In the middle and long term, prognostic consensus about climate change signal a rise in CO 2 concentration and temperature on the Earth’s surface, unexpected variations in rainfall, and more recurrent and intense weather conditions, e.g., heat waves, drought and flooding events ( Mittler and Blumwald, 2010 ; IPCC, 2014 ). These brief episodes bring plants beyond their capacity of adaptation; decreasing crop and tree yield ( Ciais et al., 2005 ; Zinta et al., 2014 ).

Here we will not discuss plants capacity of adaptation to novel environmental conditions when considering large scales and long-term periods. Ecosystems are being affected by climate change at all levels (terrestrial, freshwater, and marine), and it was already reported that species are under evolutionary adaptation to human-caused climate change (for a review see Scheffers et al., 2016 ). Migration and plasticity are two biological mechanisms to cope with these changes. Data indicate that each population of a species has limited tolerance to sharp climate variations, and they could migrate to find more favorable environments. Habitat fragmentation limits plant movement, being other big threat for adaptation ( Stockwell et al., 2003 ; Leimu et al., 2010 ). Despite the fact that individual plants are immobile, plant populations move when seeds are dispersed, resulting in differences in the general distribution of the species ( Corlett and Westcott, 2013 ). In this sense, anthropogenic activities also contribute to seed dispersal.

Plasticity is a characteristic related to phenology and phenotype. Phenology is the timing of phases occurrence in the life cycle, and phenotypic plasticity is the range of phenotypes that a single genotype may express depending on its environment ( Nicotra et al., 2010 ). Plasticity is adaptive when the phenotype changes occur in a direction favored by selection in the new environment.

Climate Change and ROS

Reactive Oxygen Species (ROS) are continuously generated by plants under normal conditions. However, they are increased in response to different abiotic stresses. One of the most important effects of climate change-related stresses at the molecular level is the increase of ROS inside the cells ( Farnese et al., 2016 ). Among ROS, the most studied are superoxide anion ( O 2 •– ), H 2 O 2 and the hydroxyl radical (⋅OH - ).

Reactive Oxygen Species cause damage to proteins, lipids and DNA, affecting cell integrity, morphology, physiology, and, consequently, the growth of plants ( Frohnmeyer and Staiger, 2003 ). The main sources of ROS in stress conditions are: augmented photorespiration, NADPH oxidase (NOX) activity, β-oxidation of fatty acids and disorders in the electron transport chains of mitochondrias and chloroplasts ( Apel and Hirt, 2004 ; AbdElgawad et al., 2015 ). Hence, higher plants have evolved in the presence of ROS and have acquired pathways to protect themselves from its toxicity. Plant antioxidant system (AS) includes the activity of ROS detoxifying enzymes [e.g., superoxide dismutase (SOD), ascorbate peroxidase (APX), catalase (CAT), glutathione peroxidase (GPX), and peroxiredoxin (PRX)], as well as antioxidant molecules such as ascorbic acid (ASC) and glutathione (GSH) that are present in almost all subcellular compartments (reviewed by Choudhury et al., 2017 ).

In this context, plants have also developed a tight interaction between ROS and NO as a mechanism to reduce the deleterious consequences of these ROS-induced oxidative injuries. NO orchestrates a wide range of mechanisms leading to the preservation of redox homeostasis in plants. Consequently, NO at low concentration is considered a broad-spectrum anti-stress molecule ( Lamattina et al., 2003 ; Tossi et al., 2009 ; Correa-Aragunde et al., 2015 ). Figure ​ Figure1 1 shows the relationship among the different GHG and their impact on plants.

An external file that holds a picture, illustration, etc.
Object name is fpls-09-00273-g001.jpg

Simplified scheme showing greenhouse gasses (GHG) and their effects on plants. GHG (H 2 O vapor, clouds, CO 2 , CH 4 , N 2 O, and NO) have both natural and anthropogenic origin, contributing to greenhouse effect. Short-term effects of GHG increase is mainly CO 2 rise, that activates photosynthesis (PS) and inhibits stomatal opening (SO). Long-term effects of GHG increase are extreme climate changes such as floods, droughts, heat. All of them induce the generation of reactive oxygen species (ROS) and oxidative stress in plants. Nitric oxide (NO) could alleviate oxidative stress by scavenging ROS and/or regulating the antioxidant system (AS). GHG and volatile organic compounds (VOC) react in presence of sunlight (E#) to give tropospheric O 3 . Although tropospheric O 3 is prejudicial for life, stratospheric O 3 is beneficial, because filters harmful UV-B radiation. The size of arrows are representative of the GHG concentration.

CO 2 and NO Contribute to Regulate Redox Homeostasis in Plants

Co 2 increasing: advantages and disadvantages.

Increased CO 2 was suggested to have a “fertilization” effect, because crops would increase their photosynthesis and stomatal conductance in response to elevated CO 2 . This belief was supported by studies performed in greenhouses, laboratory controlled-environment chambers, and transparent field chambers, where emitted CO 2 may be held back and readily controlled ( Drake et al., 1997 ; Markelz et al., 2014 ). However, more realistic results, obtained by Free-Air Concentration Enrichment (FACE) technology, suggest that the fertilization response due to CO 2 increase is probably dependent on genetic and environmental factors, and the duration of the study ( Smith and Dukes, 2013 ). An extensive review of the literature in this field made by Xu et al. (2015) concluded that augmented CO 2 normally increases photosynthesis in C3 species such as rice, soybean and wheat. On the other hand, they pointed out that a negative feedback of photosynthesis could take place in augmented CO 2 , as a result of overload of chemical and reactive generated substrates, leading to an imbalance in the sink:source carbon ratio. Moreover, the energetic cost of carbohydrate exportation increases in elevated CO 2 level.

The most important photosynthetic enzyme is the ribulose-1,5-bisphosphate carboxylase-oxygenase (RuBisCO). Rubisco is located in mesophyll cells of C3 plants, in direct contact with the intercellular air space linked to the atmosphere by epidermal stomatal pores. Photosynthesis increases at high CO 2 , because Rubisco is not CO 2 saturated and CO 2 inhibits the oxygenation reactions and photorespiration ( Long et al., 2006 ). However, long-term high concentration of CO 2 may down regulate Rubisco activity because ribulose-1,5-bisphosphate is not regenerated. Hexokinase (HXK), a sensor of extreme photosynthate, may participate in the down regulation of Rubisco concentration ( Xu et al., 2015 ). Moreover, severe abiotic stresses, such as temperature and drought, may restrain Rubisco carboxylation and foster oxygenation ( Xu et al., 2015 ).

In C4 crops, such as maize and sorghum, the elevated concentration of CO 2 inside the bundle sheath cells could prevent a large increase of Rubisco activity at higher atmospheric CO 2 and, thereby, photosynthetic activity is not augmented. However, at high CO 2 levels, the water status of C4 plants under drought conditions is improved, increasing photosynthesis and biomass accumulation ( Long et al., 2006 ; Mittler and Blumwald, 2010 ). That envisages potential advantages for the C4 species in future climatic change scenarios, particularly in arid and semiarid areas.

In addition, high CO 2 has the benefit of reducing stomatal conductance, decreasing 10% evapotranspiration in both C3 and C4 plants. Simultaneously, the cooling decreased resulting from reduced transpiration causes elevated canopy temperatures of around 0.7°C for most crops. Biomass and yield rise due to high CO 2 in all C3 plants, but not in C4 plants exception made when water is a restraint. Yields of C3 grain crops jump around 19% on average at high CO 2 ( Kimball, 2016 ).

Some reports analyze the contribution of CO 2 in the responses of plants to the combination of multiple stresses. For Arabidopsis thaliana , the combination of heat and drought induces photosynthesis inhibition of 62% under ambient CO 2 , but the drop in photosynthesis is just 40% at high CO 2 . Moreover, the protein oxidation increases significantly during a heat wave and drought, and this effect is repressed by increased CO 2 . Photorespiration is also reduced by high CO 2 ( Zinta et al., 2014 ).

Studying grasses ( Lolium perenne, Poa pratensis ) and legumes ( Medicago lupulina, Lotus corniculatus ) exposed to drought, high temperature and augmented CO 2 , AbdElgawad et al. (2015) demonstrated that drought suppresses plant growth, photosynthesis and stomatal conductance, and promotes in all species the synthesis of osmolytes and antioxidants. Instead, oxidative damage is more markedly observed in legumes than in grasses. In general, warming amplifies drought consequences. In contrast, augmented CO 2 diminishes stress impact. Reduction in photosynthesis and chlorophyll, as a result of drought and elevated temperature, were avoided by high CO 2 in the grasses. Noxious effects of oxidative stress, i.e., lipid peroxidation, are phased down in all species by augmented CO 2 . Normally, a reduced impact of oxidative stress is due to decreased photorespiration and diminished NOX activity. In legumes, a rise in levels of antioxidant molecules (flavonoids and tocopherols) contribute as well to the stress mitigation caused by augmented CO 2 . The authors draw the conclusion that these different responses point at an unequal future impact of climate change on the production of agricultural-scale legumes and grass crops.

Kumari et al. (2015) assessed the impact of various levels of CO 2 , ambient (382 ppm) and augmented (570 ppm), and O 3 , ambient (50 ppb) and augmented (70 ppb) on the potato physiological and biochemical responses ( Solanum tuberosum ). They observed that augmented CO 2 cut down O 3 uptake, enhanced carbon assimilation, and curbed oxidative stress. Elevated CO 2 also mitigated the noxious effect of high O 3 on photosynthesis.

Although some molecular mechanisms underpinning CO 2 actions are unknown, the results presented highlight the importance of CO 2 as a regulator that mitigates the potential climate change-induced deleterious consequences in plants. Recent reports suggest that some CO 2 -associated responses may be mediated by NO.

Du et al. (2016) determined that 800 μmol.mol -1 of CO 2 increased the NO concentration in Arabidopsis leaves, through a mechanism related to nitrate availability. Moreover, NO increase, as a consequence of high CO 2 levels, was reported as a general procedure to improve iron (Fe) nutrition in response to Fe deficiency in tomato roots ( Jin et al., 2009 ).

The gas exchange between the atmosphere and plants is mainly regulated by stomata. But structure and physiology of stomata are also influenced by gasses ( García-Mata and Lamattina, 2013 ). Elevated CO 2 regulate stomatal density and conductance. Moreover, there is increasing evidence that this response is modified by interaction of CO 2 with other environmental factors ( Xu et al., 2016 ; Yan et al., 2017 ). Wang et al. (2015) reported that 800 μmol.mol -1 of CO 2 increases the NO concentration in A. thaliana guard cells, inducing stomatal closure. Both NR and NO synthase (NOS)-like activities are necessary for CO 2 -induced NO accumulation. Comprehensive pharmacological and genetic results obtained in Arabidopsis by Chater et al. (2015) , show that when CO 2 concentration is around 700–1000 ppm, stomatal density and closure are reduced. They also illustrate that those elements necessary for this process are: activation of both ABA biosynthesis genes and the PYR/RCAR ABA receptor, and ROS increase. However, Shi et al. (2015) provide genetic and pharmacological evidence that high CO 2 concentration induces stomatal closure by an ABA-independent mechanism in tomato. They show that 800 μmol.mol -1 of CO 2 increase the expression of the protein kinase OPEN STOMATA 1 (OST1), NOX, and nitrate reductase (NR) genes. They also show that the sequential production of NOX-dependent H 2 O 2 and NR-produced NO are mainly dependent of OST1, and are involved in the CO 2 -induced stomatal closure.

In ABA-dependent mechanisms, ABA is increased by CO 2. The binding of ABA to its receptor (PYR/RCAR) inactivates PP2C, activating OST1. In ABA-independent mechanism, OST1 will be transcriptionally induced by CO 2 . Once activated, OST1 along with Ca 2 + , activates NOX, increasing ROS ( Kim et al., 2010 ). The rise of guard cells ROS enhances NO, cytosolic free Ca 2 + , and pH ( Song et al., 2014 ; Xie et al., 2014 ). ROS and NO release Ca 2 + from internal reservoirs, or influx external Ca 2 + through plasma membrane Ca 2 + in channels. Cytosolic free Ca 2 + inactivate inward K + channels (K + in ) to prevent K + uptake and activate outward K + channels (K + out ) and Cl - (anion) channels (Cl - ) at the plasma membrane ( Blatt, 2000 ; García-Mata et al., 2003 ). Ca 2 + also activates slow anion channel homolog 3 (SLAH3), slow anion channel-associated 1 (SLAC1) and aluminum activated malate transporters (ALMT) ( Roelfsema et al., 2012 ). The consequence of the regulation of cation/anion channels is the net efflux of K + /Cl - /malate and influx of Ca 2 + , making guard cells lose turgor by water outlet, causing stomatal closure.

All together, the results discussed here suggest that CO 2 -induced NO increase is a common plant physiological response to oxidative stresses. Figure ​ Figure2 2 shows the importance of CO 2 and NO in these processes.

An external file that holds a picture, illustration, etc.
Object name is fpls-09-00273-g002.jpg

Interplay between CO 2 and NO in plant redox physiology: CO 2 enters to the leaves by stomata. Once in mesophyll cells, CO 2 increase photosynthesis (PS) through the CO 2 -unsaturated Rubisco activity. When plants are in stress environments, ROS could be augmented by Rubisco-induced photorespiration and NADPH oxidase (NOX) activities. NOX- induced O 2 •– , in the apoplast is immediately transformed to H 2 O 2 by the superoxide dismutase (SOD). Plasma membrane is permeable to H 2 O 2 . CO 2 moderates oxidative stress in mesophyll cells by inhibiting both Rubisco photorespiration (PR) and NOX activities. Besides, NO is induced by CO 2 and ROS, alleviating the consequences of oxidative stress by scavenging ROS and activating or inhibiting the antioxidant system (AS). In guard cells, CO 2 increases the expression and activity of OPEN STOMATA 1 (OST1), in both ABA-dependent and independent mechanisms. OST1 activates NOX, producing ROS and consequently NO increase by nitrate reductase (NR), and NOS-like activities. NO prevents ROS increase by direct scavenging, and inhibiting NOX. NO-dependent Ca 2 + regulated ion channels induces stomatal closure, modulating O 3 and CO 2 uptake, decreasing evapotranspiration, and rising leaf temperature.

Abiotic Stress, ROS Generation, and Redox Balance: The Key Role of NO

Reactive oxygen species are generated in apoplast, plasma membrane, chloroplasts, mitochondria, and peroxisomes ( Farnese et al., 2016 ). It was proposed that each stress produces its own “ROS signature” ( Choudhury et al., 2017 ). For instance, drought may reduce the activity of Rubisco, decreasing CO 2 fixation and NADP+ regeneration by the Calvin cycle. As a consequence, chloroplast electron transport is altered, generating ROS by electron leakage to O 2 ( Carvalho, 2008 ). In drought stress, ROS increase is produced by NOX activity ( Farnese et al., 2016 ). In flooding, ROS generation is an ethylene-promoted process that involves calcium (Ca 2+ ) flux, and NOX activity ( Voesenek and Bailey-Serres, 2015 ).

In heat stress, a NOX-dependent transient ROS rise is an early event ( Königshofer et al., 2008 ). Then, endogenous ROS are sensed through histidine kinases, and an Arabidopsis heat stress factor (HsfA4a) appears to sense exogenous ROS. As a result, the MAPK signal pathway is activated ( Qu et al., 2013 ). Moreover, functional decrease in photosynthetic light reaction induces ROS concentration by high electron leakage from the thylakoid membrane ( Hasanuzzaman et al., 2013 ). In this process, O 2 is the acceptor, generating O 2 •– .

Thus, individual stresses or their different combinations may produce particular “ROS signatures.” Besides their deleterious effects, ROS are recognized as a signal in the plant reaction to biotic and abiotic stressors. ROS may induce programed cell death (PCD) to avoid pathogen spread ( Mur et al., 2008 ), trigger a systemic defense response signal ( Dubiella et al., 2013 ), or avoid the chloroplast antenna overloading by electrons divert ( Choudhury et al., 2017 ).

Whatever the origin and function, ROS concentration must be adequately regulated to avoid excessive concentration and consequent cellular damages. Depending on NO and ROS concentrations, NO has the dual capacity to activate or inhibit the ROS production, and is a key molecule for keeping cellular redox homeostasis under control ( Beligni and Lamattina, 1999a ; Correa-Aragunde et al., 2015 ). NO has a direct ROS-scavenging activity because it holds an unpaired electron, reaching elevated reactivity with O 2 , O 2 •– , and redox active metals. NO can mitigate OH formation by scavenging either Fe or O 2 •– ( Lamattina et al., 2003 ). However, NO reacting with ROS (mainly O 2 •– ) may generate reactive nitrogen species (RNS). An excess of RNS originates a nitrosative stress ( Corpas et al., 2011 ). To avoid the toxicity of nitrosative stress, NO is stored as GSNO in the cell.

GSH as a Redox Buffer. GSNO as NO Reservoir. SNO and S-Nitrosylation

Glutathione (GSH) is a small peptide with the sequence γ-l-glutamyl-l-cysteinyl-glycine that has a cell redox homeostatic impact in most plant tissues. It is a soluble small thiol considered a non-enzymatic antioxidant. It exists in the reduced (GSH) or oxidized state (GSSG), in which two GSH molecules are joined by a disulfide bond ( Rouhier et al., 2008 ). GSH alleviates oxidative damages in plants generated by abiotic stresses, including salinity, drought, higher, low temperature, and heavy metals. GSH is precursor of phytochelatins, polymers that chelate toxic metals and transport them to the vacuole ( Grill et al., 1989 ). Studies shown that GSH contributes to tolerate nickel, cadmium, zinc, mercury, aluminum and arsenate heavy metals in plants ( Asgher et al., 2017 ). Moreover, GSH has a role in the detoxification of ROS both directly, interacting with them, or indirectly, participating of enzymatic pathways. GSH is involved in glutathionylation, a posttranslational modification that causes a mixed disulfide bond between a Cys residue and GSH.

GSH can be oxidized to GSSG by H 2 O 2 and can react with NO to form the nitrosoglutathione (GSNO) derivative. GSNO is an intracellular NO reservoir. It is also a vehicle of NO throughout the cell and organs, spreading NO biological function. GSNO is the largest low-molecular-mass S-nitrosothiol (SNO) in plant cells ( Corpas et al., 2013 ). GSNO metabolism and its reaction with other molecules involve S-nitrosylation and S-transnitrosation which consist of the binding of a NO molecule to a cysteine residue in proteins. Thioredoxin produces protein denitrosylation ( Correa-Aragunde et al., 2013 ). GSNO could be decomposed by the GSNO reductase (GSNOR) to GSSG which, in turn, is reduced to GSH by glutathione reductase (GR).

Glutathione also participates in the GSH/ASC cycle, a series of enzymatic reactions that degrade H 2 O 2 . APX degrades H 2 O 2 using ASC, the other major antioxidant in plants, as cofactor. The oxidized ASC is reduced by monodehydroascorbate reductase (MDHAR) in an NAD(P)H-dependent manner and by dehydroascorbate reductase (DHAR) employing GSH as electron donor. The resulting GSSG is reduced in turn to GSH by GR ( Foyer and Noctor, 2011 ).

Different Effects of NO in the Regulation of Antioxidant Enzymes

The application of NO donors alleviates oxidative stress in plants challenged to abiotic and/or biotic stresses ( Laxalt et al., 1997 ; Beligni and Lamattina, 1999b , 2002 ; Shi et al., 2007 ; Xue et al., 2007 ; Leitner et al., 2009 ).

Besides the direct ROS-scavenging activity of NO, its beneficial effect is exerted by the regulation of the antioxidant enzymes activity that controls toxic levels of ROS and RNS ( Uchida et al., 2002 ; Shi et al., 2005 ; Song et al., 2006 ; Romero-Puertas et al., 2007 ; Bai et al., 2011 ). NO can modulate cell redox balance in plants through the regulation of gene expression, posttranslational modification or by its binding to the heme prosthetic group of some antioxidant enzymes.

SOD catalyzes the dismutation of stress-generated O 2 •– in one of two less harmful species: either molecular oxygen (O 2 ) or hydrogen peroxide (H 2 O 2 ). APX and CAT are the most important enzymes degrading H 2 O 2 in plants. They transform H 2 O 2 to H 2 O and O 2 . APX isoforms are primarily found in the cytosol and chloroplasts, while the CAT isoforms are found in peroxisomes. APX has strong affinity for H 2 O 2 and uses ASC as an electron donor. In contrast, CAT removes H 2 O 2 generated in the peroxisomal respiratory pathway without the need to reduce power. Even though CAT affinity for H 2 O 2 is low, its elevated rate of reaction offers an effective way to detoxify H 2 O 2 inside the cell. PRX may reduce both hydroperoxide and peroxynitrite.

Many reports on different plant species demonstrate that NO induces the transcription and activity of antioxidative enzymes in response to oxidative stress. The tolerance to drought and salt-induced oxidative stress in tobacco is related to the ABA-triggered production of H 2 O 2 and NO. In turn, they induce transcripts and activities of SOD, CAT, APX, and GR ( Zhang et al., 2009 ). UV-B-produced oxidative stress in Glycine max was alleviated by NO donors, which induced transcription and activities of SOD, CAT, and APX ( Santa-Cruz et al., 2014 ). Furthermore, in bean leaves, SOD, CAT, and APX activities are increased by NO donors, and protected from the oxidative stress generated by UV-B irradiation ( Shi et al., 2005 ). Drought tolerance in bermudagrass is improved by ABA-dependent SOD and CAT activities. This effect is regulated by H 2 O 2 and NO, NO acting downstream H 2 O 2 ( Lu et al., 2009 ).

Several antioxidant enzymes have been identified as target of S-nitrosylation, resulting in a change of their biological activity ( Romero-Puertas et al., 2008 ; Bai et al., 2011 ; Fares et al., 2011 ). For instance, NO reinforces recalcitrant seed desiccation tolerance in Antiaris toxicaria by activating the ascorbate-glutathione cycle through S-nitrosylation to control H 2 O 2 accumulation. Desiccation treatment reduced the level of S-nitrosylated APX, GR, and DHAR proteins. Instead, NO gas exposure activated them by S-nitrosylation ( Bai et al., 2011 ). Furthermore, APX was S-nitrosylated at Cys32 during saline stress and biotic stress, enhancing its enzymatic activity ( Begara-Morales et al., 2014 ; Yang et al., 2015 ). In addition, auxin-induced denitrosylation of cytosolic APX provoked inhibition of its activity, followed by an increase of H 2 O 2 concentration and the consequent lateral root formation in Arabidopsis ( Correa-Aragunde et al., 2013 ). Moreover, an inhibitory impact of S-nitrosylation on APX activity was also reported during programmed cell death in Arabidopsis ( de Pinto et al., 2013 ). CAT was identified to be S-nitrosylated in a proteomic study of isolated peroxisomes ( Ortega-Galisteo et al., 2012 ). A decrease of S-nitrosylated CAT under Cd treatment was reported. In addition, in vitro experiments demonstrated a reversible inhibitory effect of APX and CAT activities by NO binding to the Fe of the heme cofactor ( Brown, 1995 ; Clark et al., 2000 ). In addition, NOXs have been involved in plant defense, development, hormone biosynthesis and signaling ( Marino et al., 2012 ). Whereas S-nitrosylation did not affect SOD activities, nitration inhibited Mn-SOD1, Fe-SOD3, and CuZn-SOD3 activity to different degrees ( Holzmeister et al., 2015 ). SOD isoforms could also regulate endogenous NO availability by competing for the common substrate, O 2 •– , and it was demonstrated that bovine SOD may release NO from GSNO ( Singh et al., 1999 ). When GSNO is decomposed by GSNOR, it produces GSSG. GSNOR is also regulated by NO. Frungillo et al. (2014) demonstrated that NO-derived from nitrate assimilation in Arabidopsis inhibited GSNOR1 by S-nitrosylation, preventing GSNO degradation. They proposed that (S)NO controls its own generation and scavenging by modulating nitrate assimilation and GSNOR1 activity. It was also shown that chilling treatment in poplar increased S-nitrosylation of NR, along with a significant decrease of its activity ( Cheng et al., 2015 ).

The dual activity of Prx, suggests a role for this enzyme both in ROS and RNS regulation. S-nitrosylation of Arabidopsis PrxIIE inhibits its peroxynitrite activity, increasing peroxynitrite-mediated tyrosine nitration ( Romero-Puertas et al., 2007 ). Pea mitochondrial PrxIIF was S-nitrosylated under salt stress, and its peroxidase activity was reduced by 5 mM GSNO ( Camejo et al., 2013 ).

An interesting study demonstrated that NO controls hypersensitive response (HR) through S-nitrosylation of NOX, inhibiting ROS synthesis. This triggers a feedback loop limiting HR ( Yun et al., 2011 ).

Other proteins related to abiotic stress response are regulated by S-nitrosylation (For a review see Fancy et al., 2017 ).

Figure ​ Figure3 3 is a simplified diagram that illustrates the main oxidative and nitrosative effects that modulate the activities of key cell components, thus maintaining cell redox balance. Note the feedback and positive-negative regulatory processes occurring in the main pathways. They involve posttranslational modifications that activate and inhibit the components involved in cell antioxidant system.

An external file that holds a picture, illustration, etc.
Object name is fpls-09-00273-g003.jpg

Molecules and mechanisms involved in NO-mediated redox balance. H 2 O 2 is generated mainly by NOX and SOD as a response to (a)biotic stress. APX and CAT are the main H 2 O 2 -degrading enzymes. NO is increased by H 2 O 2 through the induction of NR/NOS-like activities, and may scavenge ROS or induce both the transcription and activity of SOD, CAT, and APX. In parallel, NO is combined with GSH to form nitrosoglutathione GSNO. GSNO regulates many enzymatic activities by the posttranslational modification of cysteine residues through S-Nitrosylation. NOX and CAT activities are inhibited by S-nitrosylation, whereas APX is either activated or inhibited by S-nitrosylation. NO also inhibits APX by binding to heme group. GSNO is degraded by GSNOR, which could be inhibited by H 2 O 2 and S-nitrosylation.NR could be inhibited by S-nitrosylation. GR reduces GSSG to GSH, and it is activated by S-nitrosylation. Ascorbate (ASC) is a cofactor of APX. Reduced ASC is generated by MDHAR and DHAR, using GSH as electron donor. Both enzymes are inhibited by S-nitrosylation. Reactive Nitrogen Species (RNS) may be originated by NO and O 2 •– reaction. SOD regulate RNS dismutating O 2 •– . Peroxiredoxins (Prx) reduce both ROS AND RNS. RNS are degraded by PrxIIe, and H 2 O 2 by PrxIIF. Both enzymes are inhibited by S-nitrosylation. Red lines: H 2 O 2 -regulated reactions. Purple lines: NO-regulated reactions. Green lines: GSNO-regulated reactions.

Conclusions and Perspectives

The accelerating rate of climate change, together with habitat fragmentation caused by human activity, are part of the selective pressures building a new Earth’s landscape.

Climate change is a multidimensional and simultaneous variation in duration, frequency and intensity of parameters like temperature and precipitation, altering the seasons and life on the Earth. In this scenario, plant species with increased adaptive plasticity will be better equipped to tolerate changes in the frequency of extreme weather events. GHG are one of the forces driving climate change. However, CO 2 and NO may contribute to maintaining the cell redox homeostasis, regulating the amount of ROS, GSH, GSNO, and SNO.

In this manuscript, we summarize the available evidence supporting the presence of broad spectrum anti-stress molecules, as NO in plants, for coping with unprecedented changes in environmental conditions. Future research should focus in better understanding the influence of GHG on plant physiology.

Author Contributions

RC conceived the project and wrote the manuscript. MN drew figures and collaborated in writing the manuscript. NC-A and LL supervised and complemented the drafting. All the persons entitled to authorship have been named and have approved the final version of the submitted manuscript.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest. The reviewer MCR-P and handling Editor declared their shared affiliation.

Acknowledgments

We thank ANPCYT for MN fellowship. We also thank Marta Terrazo for helping with the language revision of the manuscript.

Funding. This work was supported by grants from the Consejo Nacional de Investigaciones Cientificas y Tecnicas, the Agencia Nacional de Promoción Científica y Tecnológica, and the Universidad Nacional de Mar del Plata, Argentina. NC-A, LL, and RC are permanent members of the Scientific Research career of CONICET. MN is doctoral fellow of the ANPCYT.

1 https://ozonewatch.gsfc.nasa.gov/facts/ozone.html

  • AbdElgawad H., Farfan-Vignolo E. R., de Vos D., Asard H. (2015). Elevated CO2 mitigates drought and temperature-induced oxidative stress differently in grasses and legumes. Plant Sci. 231 1–10. 10.1016/j.plantsci.2014.11.001 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Apel K., Hirt H. (2004). Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu. Rev. Plant Biol. 55 373–399. 10.1146/annurev.arplant.55.031903.141701 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Asgher M., Per T. S., Anjum S., Khan M., Masood A., Verna S., et al. (2017). “Contribution of glutathione in heavy metal stress tolerance in plants,” in Reactive Oxygen Species and Antioxidant Systems in Plants: Role and Regulation under Abiotic Stress eds Khan M., Khan N. (Singapore: Springer; ) 10.1007/978-981-10-5254-5_12 [ CrossRef ] [ Google Scholar ]
  • Bai X., Yang L., Tian M., Chen J., Shi J., Yang Y., et al. (2011). Nitric oxide enhances desiccation tolerance of recalcitrant Antiaris toxicaria seeds via protein S-nitrosylation and carbonylation. PLoS One 6 : e20714 . 10.1371/journal.pone.0020714 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Begara-Morales J. C., Sánchez-Calvo B., Chaki M., Valderrama R., Mata-Pérez C., López-Jaramillo J., et al. (2014). Dual regulation of cytosolic ascorbate peroxidase (APX) by tyrosine nitration and S -nitrosylation. J. Exp. Bot. 65 527–538. 10.1093/jxb/ert396 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Beligni M. V., Lamattina L. (1999a). Is nitric oxide toxic or protective? Trends Plant Sci. 4 299–300. 10.1016/S1360-1385(99)01451-X [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Beligni M. V., Lamattina L. (1999b). Nitric oxide protects against cellular damage produced by methylviologen herbicides in potato plants. Nitric Oxide 3 199–208. 10.1006/niox.1999.0222 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Beligni M. V., Lamattina L. (2002). Nitric oxide interferes with plant photo-oxidative stress by detoxifying reactive oxygen species. Plant Cell Environ. 25 737–748. 10.1046/j.1365-3040.2002.00857.x [ CrossRef ] [ Google Scholar ]
  • Blatt M. R. (2000). Cellular signaling and volume control in stomatal movements in plants. Annu. Rev. Cell Dev. Biol. 16 221–241. 10.1146/annurev.cellbio.16.1.221 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Brown G. C. (1995). Reversible binding and inhibition of catalase by nitric oxide. Eur. J. Biochem. 232 188–191. 10.1111/j.1432-1033.1995.tb20798.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Camejo D., Romero-Puertas M. D. C., Rodríguez-Serrano M., Sandalio L. M., Lázaro J. J., Jiménez A., et al. (2013). Salinity-induced changes in S-nitrosylation of pea mitochondrial proteins. J. Proteomics 79 87–99. 10.1016/j.jprot.2012.12.003 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Carvalho M. H. C. (2008). Drought stress and reactive oxygen species. Plant Signal. Behav. 3 156–165. 10.4161/psb.3.3.5536 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Chamizo-Ampudia A., Sanz-Luque E., Llamas A., Ocana-Calahorro F., Mariscal V., Carreras A., et al. (2016). A dual system formed by the ARC and NR molybdoenzymes mediates nitrite-dependent NO production in Chlamydomonas . Plant Cell Environ. 39 2097–2107. 10.1111/pce.12739 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Cheng T., Chen J., Ef A. A., Wang P., Wang G., Hu X., et al. (2015). Quantitative proteomics analysis reveals that S-nitrosoglutathione reductase (GSNOR) and nitric oxide signaling enhance poplar defense against chilling stress. Planta 242 1361–1390. 10.1007/s00425-015-2374-5 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Choudhury F. K., Rivero R. M., Blumwald E., Mittler R. (2017). Reactive oxygen species, abiotic stress and stress combination. Plant J. 90 856–867. 10.1111/tpj.13299 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ciais P., Reichstein M., Viovy N., Granier A., Ogée J., Allard V., et al. (2005). Europe-wide reduction in primary productivity caused by the heat and drought in 2003. Nature 437 529–533. 10.1038/nature03972 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Clark D., Durner J., Navarre D. A., Klessig D. F. (2000). Nitric oxide inhibition of tobacco catalase and ascorbate peroxidase. Mol. Plant. Microbe. Interact. 13 1380–1384. 10.1006/niox.1999.0222 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Corlett R. T., Westcott D. A. (2013). Will plant movements keep up with climate change? Trends Ecol. Evol. 28 482–488. 10.1016/j.tree.2013.04.003 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Corpas F. J., Alché J. D., Barroso J. B. (2013). Current overview of S-nitrosoglutathione (GSNO) in higher plants. Front. Plant Sci. 4 : 126 . 10.3389/fpls.2013.00126 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Corpas F. J., Leterrier M., Valderrama R., Airaki M., Chaki M., Palma J. M., et al. (2011). Nitric oxide imbalance provokes a nitrosative response in plants under abiotic stress. Plant Sci. 181 604–611. 10.1016/j.plantsci.2011.04.005 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Correa-Aragunde N., Foresi N., Delledonne M., Lamattina L. (2013). Auxin induces redox regulation of ascorbate peroxidase 1 activity by S-nitrosylation/denitrosylation balance resulting in changes of root growth pattern in Arabidopsis . J. Exp. Bot. 64 3339–3349. 10.1093/jxb/ert172 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Correa-Aragunde N., Foresi N., Lamattina L. (2015). Nitric oxide is a ubiquitous signal for maintaining redox balance in plant cells: regulation of ascorbate peroxidase as a case study. J. Exp. Bot. 66 2913–2921. 10.1093/jxb/erv073 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Chater C., Peng K., Movahedi M., Dunn J. A., Walker H. J., Liang Y. K., et al. (2015). Elevated CO 2 -induced responses in stomata require ABA and ABA signaling. Curr. Biol. 25 2709–2716. 10.1016/j.cub.2015.09.013 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • de Pinto M. C., Locato V., Sgobba A., Romero-Puertas M. D. C., Gadaleta C., Delledonne M., et al. (2013). S-Nitrosylation of ascorbate peroxidase is part of programmed cell death signaling in tobacco bright yellow-2 cells. Plant Physiol. 163 1766–1775. 10.1093/jxb/ert172 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Di Dato V., Musacchia F., Petrosino G., Patil S., Montresor M., Sanges R., et al. (2015). Transcriptome sequencing of three Pseudo-nitzschia species reveals comparable gene sets and the presence of nitric oxide synthase genes in diatoms. Sci. Rep. 5 : 12329 . 10.1038/srep12329 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Drake B. G., Gonzalez-Meler M. A., Long S. P. (1997). More efficient plants: a consequence of rising atmospheric CO 2 ? Annu. Rev. Plant Physiol. Plant Mol. Biol. 48 609–639. 10.1146/annurev.arplant.48.1.609 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Du S., Zhang R., Zhang P., Liu H., Yan M., Chen N., et al. (2016). Elevated CO 2 -induced production of nitric oxide (NO) by NO synthase differentially affects nitrate reductase activity in Arabidopsis plants under different nitrate supplies. J. Exp. Bot. 67 893–904. 10.1093/jxb/erv506 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dubiella U., Seybold H., Durian G., Komander E., Lassig R., Witte C. P., et al. (2013). Calcium-dependent protein kinase/NADPH oxidase activation circuit is required for rapid defense signal propagation. Proc. Natl. Acad. Sci. U.S.A. 110 8744–8749. 10.1073/pnas.1221294110 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fancy N. N., Bahlmann A. K., Loake G. J. (2017). Nitric oxide function in plant abiotic stress. Plant Cell Environ. 40 462–472. 10.1111/pce.12707 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Fares A., Rossignol M., Peltier J. (2011). Proteomics investigation of endogenous S-nitrosylation in Arabidopsis . Biochem. Biophys. Res. Commun. 416 331–336. 10.1038/srep12329 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Farnese F. S., Menezes-Silva P. E., Gusman G. S., Oliveira J. A. (2016). When bad guys become good ones: the key role of reactive oxygen species and Nitric Oxide in the plant responses to abiotic stress. Front. Plant Sci. 7 : 471 . 10.3389/fpls.2016.00471 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Foresi N., Correa-Aragunde N., Parisi G., Calo G., Salerno G., Lamattina L. (2010). Characterization of a nitric oxide synthase from the plant kingdom: NO generation from the green alga Ostreococcus tauri is light irradiance and growth phase dependent. Plant Cell 22 3816–3830. 10.1105/tpc.109.073510 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Foyer C. H., Noctor G. (2011). Ascorbate and glutathione: the heart of the redox hub. Plant Physiol. 155 2–18. 10.1038/srep12329 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Frohlich A., Durner J. (2011). The hunt for plant nitric oxide synthase (NOS): Is one really needed? Plant Sci. 181 401–404. 10.1016/j.plantsci.2011.07.014 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Frohnmeyer H., Staiger D. (2003). Ultraviolet-B-radiation-mediated responses in plants. Balancing damage and protection. Plant Physiol. 133 1420–1428. 10.1104/pp.103.030049 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Frungillo L., Skelly M. J., Loake G. J., Spoel S. H., Salgado I. (2014). S-nitrosothiols regulate nitric oxide production and storage in plants through the nitrogen assimilation pathway. Nat. Commun. 5 : 5401 . 10.1038/ncomms6401 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • García-Mata C., Gay R., Sokolovski S., Hills A., Lamattina L., Blatt M. R. (2003). Nitric oxide regulates K + and Cl-channels in guard cells through a subset of abscisic acid-evoked signaling pathways. Proc. Natl. Acad. Sci. U.S.A. 100 11116–11121. 10.1073/pnas.1434381100 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • García-Mata C., Lamattina L. (2013). Gasotransmitters are emerging as new guard cell signaling molecules and regulators of leaf gas exchange. Plant Sci. 201-202 66–73. 10.3389/fpls.2016.00277 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gregg J., Jones C., Dawson T. (2003). Urbanization effects on tree growth in the vicinity of New York City. Nature 424 183–187. 10.1038/nature01728 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Grill E., Löffler S., Winnacker E. L., Zenk M. H. (1989). Phytochelatins, the heavy-metal-binding peptides of plants, are synthesized from glutathione by a specific gamma-glutamylcysteine dipeptidyl transpeptidase (phytochelatin synthase). Proc. Natl. Acad. Sci. U.S.A. 86 6838–6842. 10.1038/srep12329 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Gupta K. J., Igamberdiev A. U. (2016). Reactive nitrogen species in mitochondria and their implications in plant energy status and hypoxic stress tolerance. Front. Plant Sci. 7 : 369 . 10.3389/fpls.2016.00369 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hall S., Huber D., Grimm N. (2008). Soil N 2 O and NO emissions from an arid, urban ecosystem. J. Geophys. Res. 113 : G01016 10.1029/2007JG000523 [ CrossRef ] [ Google Scholar ]
  • Hasanuzzaman M., Nahar K., Alam M. M., Roychowdhury R., Fujita M. (2013). Physiological, biochemical, and molecular mechanisms of heat stress tolerance in plants. Int. J. Mol. Sci. 14 9643–9684. 10.3390/ijms14059643 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Held I., Souden B. (2000). Water vapor feedback and global warming. Annu. Rev. Energy Environ. 25 441–475. 10.1146/annurev.energy.25.1.441 [ CrossRef ] [ Google Scholar ]
  • Holzmeister C., Gaupels F., Geerlof A., Sarioglu H., Sattler M., Durner J., et al. (2015). Differential inhibition of Arabidopsis superoxide dismutases by peroxynitrite-mediated tyrosine nitration. J. Exp. Bot. 66 989–999. 10.1093/jxb/eru458 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • IPCC (2014). Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change eds Edenhofer O. R., Pichs-Madruga Y., Sokona E., Farahani S., Kadner K., Seyboth A., et al. Cambridge: Cambridge University Press. [ Google Scholar ]
  • Jeandroz S., Wipf D., Stuehr D. J., Lamattina L., Melkonian M., Tian Z., et al. (2016). Occurrence, structure, and evolution of nitric oxide synthase-like proteins in the plant kingdom. Sci. Signal. 9 : re2 . 10.1126/scisignal.aad4403 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jhun I., Coull B. A., Zanobetti A., Koutrakis P. (2015). The impact of nitrogen oxides concentration decreases on ozone trends in the USA. Air Qual. Atmos. Health 8 283–292. 10.1088/1748-9326/10/8/084009 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jin C. W., Du S. T., Chen W. W., Li G. X., Zhang Y. S., Zheng S. J. (2009). Elevated carbon dioxide improves plants iron nutrition through enhancing the iron-deficiency-induced responses under iron-limited conditions in tomato. Plant Physiol. 150 272–280. 10.1104/pp.109.136721 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Karl T. R., Trenberth K. E. (2003). Modern global climate change. Science 302 1719–1723. 10.1126/science.1090228 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kiehl J., Trenberth K. (1997). Earth’s annual global mean energy budget. Bull. Amer. Meteor. Soc. 78 197–208. [ Google Scholar ]
  • Kim T., Böhmer M., Hu H., Nishimura N., Schroeder J. I. (2010). Guard cell signal transduction network: advances in understanding abscisic acid, CO 2 , and Ca 2+ signaling. Annu. Rev. Plant Biol. 61 561–591. 10.1146/annurev-arplant-042809-112226 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kimball B. (2016). Crop responses to elevated CO 2 and interactions with H 2 O, N, and temperature. Curr. Opin. Plant Biol. 31 36–43. 10.1016/j.pbi.2016.03.006 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Königshofer H., Tromballa H. W., Löppert H. G. (2008). Early events in signaling high-temperature stress in tobacco BY2 cells involve alterations in membrane fluidity and enhanced hydrogen peroxide production. Plant Cell Environ. 31 1771–1780. 10.1111/j.1365-3040.2008.01880.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kumar A., Castellano I., Patti F. P., Palumbo A., Buia M. C. (2015). Nitric oxide in marine photosynthetic organisms. Nitric Oxide 47 34–39. 10.1038/srep12329 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Kumari S., Agrawal M., Singh A. (2015). Effects of ambient and elevated CO 2 and ozone on physiological characteristics, antioxidative defense system and metabolites of potato in relation to ozone flux. Environ. Exp. Bot. 109 276–287. 10.1016/j.envexpbot.2014.06.015 [ CrossRef ] [ Google Scholar ]
  • Lamattina L., García-Mata C., Graziano M., Pagnussat G. (2003). Nitric oxide: the versatility of an extensive signal molecule. Annu. Rev. Plant Biol. 54 109–136. 10.1146/annurev.arplant.54.031902.134752 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Laxalt A. M., Beligni M. V., Lamattina L. (1997). Nitric oxide preserves the level of chlorophyll in potato leaves infected by Phytophthora infestans. Eur. J. Plant Pathol. 103 643–651. 10.1023/A:1008604410875 [ CrossRef ] [ Google Scholar ]
  • Le Treut H., Somerville R., Cubasch U., Ding Y., Mauritzen C., Mokssit A., et al. (2007). “Historical overview of climate change science,” in Climate change 2007: The physical science basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change eds Solomon S., Qin D., Manning M., Chen Z., Marquis M., Averyt K. B., et al. (Cambridge: Cambridge University Press; ). [ Google Scholar ]
  • Leimu R., Vergeer P., Angeloni F., Ouborg N. (2010). Habitat fragmentation, climate change, and inbreeding in plants. Ann. N. Y. Acad. Sci. 1195 84–98. 10.1111/j.1749-6632.2010.05450.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Leitner M., Vandelle E., Gaupels F., Bellin D., Delledonne M. (2009). NO signals in the haze: nitric oxide signalling in plant defence. Curr. Opin. Plant Biol. 12 451–458. 10.1016/j.pbi.2009.05.012 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Long S., Ainsworth E., Leakey A., Nösberger J., Ort D. (2006). Food for thought: lower-than-expected crop yield stimulation with rising CO 2 concentrations. Science 312 1918–1921. 10.1126/science.1114722 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Long S., Ainsworth E., Rogers A., Ort D. (2004). Rising atmospheric carbon dioxide: plants FACE the future. Annu. Rev. Plant Biol. 55 591–628. 10.1146/annurev.arplant.55.031903.141610 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Lu S., Su W., Li H., Guo Z. (2009). Abscisic acid improves drought tolerance of triploid bermudagrass and involves H 2 O 2 - and NO-induced antioxidant enzyme activities. Plant Physiol. Biochem. 7 132–138. 10.1016/j.plaphy.2008.10.006 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Marino D., Dunand C., Puppo A., Pauly N. (2012). A burst of plant NADPH oxidases. Trends Plant Sci. 17 9–15. 10.1016/j.tplants.2011.10.001 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Markelz R. J., Lai L. X., Vosseler L. N., Leakey A. D. (2014). Transcriptional reprogramming and stimulation of leaf respiration by elevated CO 2 concentration is diminished, but not eliminated, under limiting nitrogen supply. Plant Cell Environ. 37 886–988. 10.1111/pce.12205 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • McAdam E. L., Brodribb T. J., McAdam S. A. (2017). Does ozone increase ABA levels by non-enzymatic synthesis causing stomata to close? Plant Cell Environ. 40 741–747. 10.1111/pce.12893 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Medinets S., Skiba U., Rennenberg H., Butterbach-Bahl K. (2015). A review of soil NO transformation: associated processes and possible physiological significance on organisms. Soil Biol. Biochem. 80 92–117. 10.1016/j.soilbio.2014.09.025 [ CrossRef ] [ Google Scholar ]
  • Meyer C., Lea U. S., Provan F., Kaiser W. M., Lillo C. (2005). Is nitrate reductase a major player in the plant NO (nitric oxide) game? Photosynth. Res. 83 181–189. 10.1016/j.tplants.2011.10.001 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mittler R., Blumwald E. (2010). Genetic engineering for modern agriculture: challenges and perspectives. Annu. Rev. Plant Biol. 61 443–462. 10.1146/annurev-arplant-042809-112116 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Mur L. A., Kenton P., Lloyd A. J., Ougham H., Prats E. (2008). The hypersensitive response; the centenary is upon us but how much do we know? J. Exp. Bot. 59 501–520. 10.1093/jxb/erm239 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Nicotra A., Atkin O., Bonser S., Davidson A., Finnegan E., Mathesius U., et al. (2010). Plant phenotypic plasticity in a changing climate. Trends Plant Sci. 15 684–692. 10.1016/j.tplants.2010.09.008 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ortega-Galisteo A. P., Rodríguez-Serrano M., Pazmiño D. M., Gupta D. K., Sandalio L. M., Romero-Puertas M. C. (2012). S-Nitrosylated proteins in pea ( Pisum sativum L.) leaf peroxisomes: changes under abiotic stress. J. Exp. Bot. 63 2089–2103. 10.1016/j.tplants.2011.10.001 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Pilegaard K. (2013). Processes regulating nitric oxide emissions from soils. Philos. Trans. R. Soc. Lond. B Biol. Sci. 368 : 20130126 . 10.1098/rstb.2013.0126 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Qu A. L., Ding Y. F., Jiang Q., Zhu C. (2013). Molecular mechanisms of the plant heat stress response. Biochem. Biophys. Res. Commun. 432 203–207. 10.1016/j.bbrc.2013.01.104 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rockel P., Strube F., Rockel A., Wildt J., Kaiser W. M. (2002). Regulation of nitric oxide (NO) production by plant nitrate reductase in vivo and in vitro. J. Exp. Bot. 53 103–110. 10.1093/jexbot/53.366.103 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Roelfsema M. R., Hedrich R., Geiger D. (2012). Anion channels: master switches of stress responses. Trends Plant Sci. 17 221–229. 10.1016/j.tplants.2012.01.009 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Romero-Puertas M. C., Campostrini N., Mattè A., Righetti P. G., Perazzolli M., Zolla L., et al. (2008). Proteomic analysis of S-nitrosylated proteins in Arabidopsis thaliana undergoing hypersensitive response. Proteomics 8 1459–1469. 10.1002/pmic.200700536 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Romero-Puertas M. C., Laxa M., Mattè A., Zaninotto F., Finkemeier I., Jones A. M. E., et al. (2007). S-nitrosylation of peroxiredoxin II E promotes peroxynitrite-mediated tyrosine nitration. Plant Cell 19 4120–4130. 10.1105/tpc.107.055061 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Rouhier N., Lemaire S. D., Jacquot J.-P. (2008). The role of glutathione in photosynthetic organisms: emerging functions for glutaredoxins and glutathionylation. Annu. Rev. Plant Biol. 59 143–166. 10.1146/annurev.arplant.59.032607.092811 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Santa-Cruz D., Pacienza N., Zilli C., Tomaro M., Balestrasse K., Yannarelli G. (2014). Nitric oxide induces specific isoforms of antioxidant enzymes in soybean leaves subjected to enhanced ultraviolet-B radiation. J. Photochem. Photobiol. B 141 202–209. 10.1016/j.jphotobiol.2014.09.019 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Scheffers B. R., De Meester L., Bridge T. C., Hoffmann A. A., Pandolfi J. M., Corlett R. T., et al. (2016). The broad footprint of climate change from genes to biomes to people. Science 354 : aaf7671 . 10.1126/science.aaf7671 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Schmidt G. A., Ruedy R. A., Miller R. L., Lacis A. A. (2010). Attribution of the present-day total greenhouse effect. J. Geophys. Res. 115 : D20106 10.1029/2010JD014287 [ CrossRef ] [ Google Scholar ]
  • Shi K., Li X., Zhang H., Zhang G., Liu Y., Zhou Y., et al. (2015). Guard cell hydrogen peroxide and nitric oxide mediate elevated CO 2 -induced stomatali movement in tomato. New Phytol. 208 342–353. 10.1111/nph.13621 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shi Q., Ding F., Wang X., Wei M. (2007). Exogenous nitric oxide protect cucumber roots against oxidative stress induced by salt stress. Plant Physiol. Biochem. 45 542–550. 10.1016/j.plaphy.2007.05.005 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Shi S., Wang G., Wang Y., Zhang L., Zhang L. (2005). Protective effect of nitric oxide against oxidative stress under ultraviolet-B radiation. Nitric Oxide 13 1–9. 10.1016/j.niox.2005.04.006 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Singh R., Hogg N., Goss S., Antholine W., Kalyanaraman B. (1999). Mechanism of superoxide dismutase/H 2 O 2 -mediated nitric oxide release from S -nitrosoglutathione–role of glutamate. Arch. Biochem. Biophys. 372 8–15. 10.1006/abbi.1999.1447 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Smith N. G., Dukes J. S. (2013). Plant respiration and photosynthesis in global-scale models: incorporating acclimation to temperature and CO 2 . Glob. Change Biol. 19 45–63. 10.1111/j.1365-2486.2012.02797.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Soden B. J., Jackson D. L., Ramaswamy V., Schwarzkopf M. D., Huang X. (2005). The radiative signature of upper tropospheric moistening. Science 310 841–844. 10.1126/science.1115602 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Song L., Ding W., Zhao M., Sun B., Zhang L. (2006). Nitric oxide protects against oxidative stress under heat stress in the calluses from two ecotypes of reed. Plant Sci. 171 449–458. 10.1016/j.plantsci.2006.05.002 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Song Y., Miao Y., Song C. P. (2014). Behind the scenes: the roles of reactive oxygen species in guard cells. New Phytol. 201 1121–1140. 10.1111/nph.12565 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Stockwell C., Hendry A., Kinnison M. (2003). Contemporary evolution meets conservation. Trends Ecol. Evol. 18 94–101. 10.1016/S0169-5347(02)00044-7 [ CrossRef ] [ Google Scholar ]
  • Tossi V., Lamattina L., Cassia R. (2009). An increase in the concentration of abscisic acid is critical for nitric oxide-mediated plant adaptive responses to UV-B irradiation. New Phytol. 181 871–879. 10.1111/j.1469-8137.2008.02722.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Uchida A., Jagendorf A. T., Hibino T., Takabe T., Takabe T. (2002). Effects of hydrogen peroxide and nitric oxide on both salt and heat stress tolerance in rice. Plant Sci. 163 515–523. 10.1016/S0168-9452(02)00159-0 [ CrossRef ] [ Google Scholar ]
  • Vainonen J. P., Kangasjärvi J. (2015). Plant signalling in acute ozone exposure. Plant Cell Environ. 38 240–252. 10.1111/pce.12273 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Voesenek L. A., Bailey-Serres J. (2015). Flood adaptive traits and processes: an overview. New Phytol. 206 57–73. 10.1111/nph.13209 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Wang H., Xiao W., Niu Y., Chai R., Jin C., Zhang Y. (2015). Elevated carbon dioxide induces stomatal closure of Arabidopsis thaliana (L.) heynh. through an increased production of nitric oxide. J. Plant Growth Regul. 34 372–380. 10.1007/s00344-014-9473-6 [ CrossRef ] [ Google Scholar ]
  • Wuebbles D. J., Hayhoe K. (2002). Atmospheric methane and global change. Earth Sci. Rev. 57 177–210. 10.1016/S0012-8252(01)00062-9 [ CrossRef ] [ Google Scholar ]
  • Xie Y., Mao Y., Zhang W., Lai D., Wang Q., Shen W. (2014). Reactive oxygen species-dependent nitric oxide production contributes to hydrogen-promoted stomatal closure in Arabidopsis. Plant Physiol. 165 759–773. 10.1104/pp.114.237925 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xu Z., Jiang Y., Jia B., Zhou G. (2016). Elevated-CO 2 response of stomata and its dependence on environmental factors. Front. Plant Sci. 7 : 657 . 10.3389/fpls.2016.00657 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xu Z., Jiang Y., Zhou G. (2015). Response and adaptation of photosynthesis, respiration, and antioxidant systems to elevated CO 2 with environmental stress in plants. Front. Plant Sci. 6 : 701 . 10.3389/fpls.2015.00701 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Xue L., Li S., Sheng H., Feng H., Xu S., An L. (2007). Nitric oxide alleviates oxidative damage induced by enhanced ultraviolet-B radiation in cyanobacterium. Curr. Microbiol. 55 294–301. 10.1007/s00284-006-0621-5 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yan W., Zhong Y., Shangguan Z. (2017). Contrasting responses of leaf stomatal characteristics to climate change: a considerable challenge to predict carbon and water cycles. Glob. Chang. Biol. 23 3781–3793. 10.1111/gcb.13654 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yang H., Mu J., Chen L., Feng J., Hu J., Li L., et al. (2015). S -Nitrosylation positively regulates ascorbate peroxidase activity during plant stress responses. Plant Physiol. 167 1604–1615. 10.1104/pp.114.255216 [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Yun B.-W., Feechan A., Yin M., Saidi N. B., Le Bihan T., Yu M., et al. (2011). S-nitrosylation of NADPH oxidase regulates cell death in plant immunity. Nature 478 264–268. 10.1038/nature10427 [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zhang Y., Tan J., Guo Z., Lu S., He S., Shu W., et al. (2009). Increased abscisic acid levels in transgenic tobacco over-expressing 9 cis-epoxycarotenoid dioxygenase influence H 2 O 2 and NO production and antioxidant defenses. Plant Cell Environ. 32 509–519. 10.1111/j.1365-3040.2009.01945.x [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Zinta G., AbdElgawad H., Domagalska M. A., Vergauwen L., Knapen D., Nijs I., et al. (2014). Physiological, biochemical, and genome-wide transcriptional analysis reveals that elevated CO 2 mitigates the impact of combined heat wave and drought stress in Arabidopsis thaliana at multiple organizational levels. Glob. Chang. Biol. 20 3670–3685. 10.1111/gcb.12626 [ PubMed ] [ CrossRef ] [ Google Scholar ]

REVIEW article

Climate change and the impact of greenhouse gasses: co 2 and no, friends and foes of plant oxidative stress.

\r\nRaúl Cassia*

  • Instituto de Investigaciones Biológicas, Facultad de Ciencias Exactas y Naturales, Universidad Nacional de Mar del Plata-Consejo Nacional de Investigaciones Científicas y Técnicas, Mar del Plata, Argentina

Here, we review information on how plants face redox imbalance caused by climate change, and focus on the role of nitric oxide (NO) in this response. Life on Earth is possible thanks to greenhouse effect. Without it, temperature on Earth’s surface would be around -19 ∘ C, instead of the current average of 14 ∘ C. Greenhouse effect is produced by greenhouse gasses (GHG) like water vapor, carbon dioxide (CO 2 ), methane (CH 4 ), nitrous oxides (N x O) and ozone (O 3 ). GHG have natural and anthropogenic origin. However, increasing GHG provokes extreme climate changes such as floods, droughts and heat, which induce reactive oxygen species (ROS) and oxidative stress in plants. The main sources of ROS in stress conditions are: augmented photorespiration, NADPH oxidase (NOX) activity, β-oxidation of fatty acids and disorders in the electron transport chains of mitochondria and chloroplasts. Plants have developed an antioxidant machinery that includes the activity of ROS detoxifying enzymes [e.g., superoxide dismutase (SOD), ascorbate peroxidase (APX), catalase (CAT), glutathione peroxidase (GPX), and peroxiredoxin (PRX)], as well as antioxidant molecules such as ascorbic acid (ASC) and glutathione (GSH) that are present in almost all subcellular compartments. CO 2 and NO help to maintain the redox equilibrium. Higher CO 2 concentrations increase the photosynthesis through the CO 2 -unsaturated Rubisco activity. But Rubisco photorespiration and NOX activities could also augment ROS production. NO regulate the ROS concentration preserving balance among ROS, GSH, GSNO, and ASC. When ROS are in huge concentration, NO induces transcription and activity of SOD, APX, and CAT. However, when ROS are necessary (e.g., for pathogen resistance), NO may inhibit APX, CAT, and NOX activity by the S-nitrosylation of cysteine residues, favoring cell death. NO also regulates GSH concentration in several ways. NO may react with GSH to form GSNO, the NO cell reservoir and main source of S-nitrosylation. GSNO could be decomposed by the GSNO reductase (GSNOR) to GSSG which, in turn, is reduced to GSH by glutathione reductase (GR). GSNOR may be also inhibited by S-nitrosylation and GR activated by NO. In conclusion, NO plays a central role in the tolerance of plants to climate change.

Introduction

Life on Earth, as it is, relies on the natural atmospheric greenhouse effect. This is the result of a process in which a planet’s atmosphere traps the sun radiation and warms the planet’s surface.

Greenhouse effect occurs in the troposphere (the lower atmosphere layer), where life and weather occur. In the absence of greenhouse effect, the average temperature on Earth’s surface is estimated around -19°C, instead of the current average of 14°C ( Le Treut et al., 2007 ). Greenhouse effect is produced by greenhouse gasses (GHG). GHG are those gaseous constituents of the atmosphere that absorb and emit radiation in the thermal infrared range ( IPCC, 2014 ). Traces of GHG, both natural and anthropogenic, are present in the troposphere. The most abundant GHG in increasing order of importance are: water vapor, carbon dioxide (CO 2 ), methane (CH 4 ), nitrous oxides (N x O) and ozone (O 3 ) ( Kiehl and Trenberth, 1997 ). GHG percentages vary daily, seasonally, and annually.

GHG Contribute Differentially to Greenhouse Effect

Water vapor.

Water is present in the troposphere both as vapor and clouds. Water vapor was reported by Tyndal in 1861 as the most important gaseous absorber of variations in infrared radiation (cited in Held and Souden, 2000 ). Further accurate calculation estimate that water vapor and clouds are responsible for 49 and 25%, respectively, of the long wave (thermal) absorption ( Schmidt et al., 2010 ). However, atmospheric lifetime of water vapor is short (days) compared to other GHG as CO 2 (years) ( IPCC, 2014 ).

Water vapor concentrations are not directly influenced by anthropogenic activity and vary regionally. However, human activity increases global temperatures and water vapor formation indirectly, amplifying the warming in a process known as water vapor feedback ( Soden et al., 2005 ).

Carbon Dioxide (CO 2 )

Carbon dioxide is responsible for 20% of the thermal absorption ( Schmidt et al., 2010 ).

Natural sources of CO 2 include organic decomposition, ocean release and respiration. Anthropogenic CO 2 sources are derived from activities such as cement manufacturing, deforestation, fossil fuels combustion such as coal, oil and natural gas, etc. Surprisingly, 24% of direct CO 2 emission comes from agriculture, forestry and other land use, and 21% comes from industry ( IPCC, 2014 ).

Atmospheric CO 2 concentrations climbed up dramatically in the past two centuries, rising from around 270 μmol.mol -1 in 1750 to present concentrations higher than 385 μmol.mol -1 ( Mittler and Blumwald, 2010 ; IPCC, 2014 ). Around 50% of cumulative anthropogenic CO 2 emissions between 1750 and 2010 have taken place since the 1970s ( IPCC, 2014 ). It is calculated that the temperature rise produced by high CO 2 concentrations, plus the water positive feedback, would increase by 3–5°C the global mean surface temperature in 2100 ( IPCC, 2014 ).

Methane (CH 4 )

Methane (CH 4 ) is the main atmospheric organic trace gas. CH 4 is the primary component of natural gas, a worldwide fuel source. Significant emissions of CH 4 result from cattle farming and agriculture, but mainly as a consequence of fossil fuel use. Concentrations of CH 4 were multiplied by two since the pre-industrial era. The present worldwide-averaged concentration is of 1.8 μmol.mol -1 ( IPCC, 2014 ).

Although its concentration represents only 0.5% that of CO 2 , concerns arise regarding a jump in CH 4 atmospheric release. Indeed, it is 30 times more powerful than CO 2 as GHG ( IPCC, 2014 ). CH 4 generates O 3 (see below), and along with carbon monoxide (CO), contributes to control the amount of OH in the troposphere ( Wuebbles and Hayhoe, 2002 ).

Nitrous Oxides (NxO)

Nitrous oxide (N 2 O) and nitric oxide (NO) are GHG. During the last century, their global emissions have rised, due mainly to human intervention ( IPCC, 2014 ). The soil emits both N 2 O and NO. N 2 O is a strong GHG, whereas NO contributes indirectly to O 3 synthesis. As GHG, N 2 O is potentially 300 times stronger than CO 2 . Once in the stratosphere, the former catalyzes the elimination of O 3 ( IPCC, 2014 ). In the atmosphere, N 2 O concentrations are climbing up due mainly to microbial activity in nitrogen (N)-rich soils related with agricultural and fertilization practices ( Hall et al., 2008 ).

Anthropogenic emissions (from combustion of fossil fuels) and biogenic emissions from soils are the main sources of NO in the atmosphere ( Medinets et al., 2015 ). In the troposphere, NO quickly oxidizes to nitrogen dioxide (NO 2 ). NO and NO 2 (termed as NO x ) may react with volatile organic compounds (VOCs) and hydroxyl, resulting in organic nitrates and nitric acid, respectively. They access ecosystems through atmospheric deposition that has an impact on the N cycle as a result of acidification or N enrichment ( Pilegaard, 2013 ).

NO Sources and Chemical Reactions in Plants

Two major pathways for NO production have been described in plants: the reductive and the oxidative pathways. The reductive pathway involves the reduction of nitrite to NO by NR under conditions such as acidic pH, anoxia, or an increase in nitrite levels ( Rockel et al., 2002 ; Meyer et al., 2005 ). NR-dependent NO formation has been involved in processes such as stomatal closure, root development, germination and immune responses. In plants, nitrite may also be reduced enzymatically by other molybdenum enzymes such as, xanthine oxidase, aldehyde oxidase, and sulfite oxidase, in animals ( Chamizo-Ampudia et al., 2016 ) or via the electron transport system in mitochondria ( Gupta and Igamberdiev, 2016 ).

The oxidative pathway produces NO through the oxidation of organic compounds such as polyamines, hydroxylamine and arginine. In animals, NOS catalyzes arginine oxidation to citrulline and NO. Many efforts were made to find the arginine-dependent NO formation in plants, as well as of plant NOS ( Frohlich and Durner, 2011 ). The identification of NOS in the green alga Ostreococcus tauri ( Foresi et al., 2010 ) led to high-throughput bioinformatic analysis in plant genomes. This study shows that NOS homologs were not present in over 1,000 genomes of higher plants analyzed, but only in few photosynthetic microorganisms, such as algae and diatoms ( Di Dato et al., 2015 ; Kumar et al., 2015 ; Jeandroz et al., 2016 ). In summary, although an arginine-dependent NO production is found in higher plants, the specific enzyme/s involved in the oxidative pathways remain elusive.

Ozone (O 3 )

Ozone (O 3 ) is mainly found in the stratosphere, but a little amount is generated in the troposphere. Stratospheric ozone (namely the ozone layer) is formed naturally by chemical reactions involving solar ultraviolet (UV) radiation and O 2 . Solar UV radiation breaks one O 2 molecule, producing two oxygen atoms (2 O). Then, each of these highly reactive atoms combines with O 2 to produce an (O 3 ) molecule. Almost 99% of the Sun’s medium-frequency UV light (from about 200 to 315 nm wavelength) is absorbed by the (O 3 ) layer. Otherwise, they could damage exposed life forms near the Earth surface 1 .

The majority of tropospheric O 3 appears when NOx, CO and VOCs, react in the presence of sunlight. However, it was reported that NOx may scavenge O 3 in urban areas ( Gregg et al., 2003 ). This dual interaction between NOx and O 3 is influenced by light, season, temperature and VOC concentration ( Jhun et al., 2015 ).

Besides, the oxidation of CH 4 by OH in the troposphere gives way to formaldehyde (CH 2 O), CO, and O 3 , in the presence of high amounts of NOx 1 .

Tropospheric O 3 is harmful to both plants and animals (including humans). O 3 affects plants in several ways. Stomata are the cells, mostly on the underside of the plant leaves, that allow CO 2 and water to diffuse into the tissue. High concentrations of O 3 cause plants to close their stomata ( McAdam et al., 2017 ), slowing down photosynthesis and plant growth. O 3 may also provoke strong oxidative stress, damaging plant cells ( Vainonen and Kangasjärvi, 2015 ).

Global Climate Change: an Integrative Balance of the Impact on Plants

Anthropogenic activity alters global climate by interfering with the flows of energy through changes in atmospheric gasses composition, more than the actual generation of heat due to energy usage ( Karl and Trenberth, 2003 ). Short-term consequences of GHG increase in plants are mainly associated with the rise in atmospheric CO 2 . Plants respond directly to elevated CO 2 increasing net photosynthesis, and decreasing stomatal opening ( Long et al., 2004 ). To a lesser extent, O 3 uptake by plants may reduce photosynthesis and induce oxidative stress. In the middle and long term, prognostic consensus about climate change signal a rise in CO 2 concentration and temperature on the Earth’s surface, unexpected variations in rainfall, and more recurrent and intense weather conditions, e.g., heat waves, drought and flooding events ( Mittler and Blumwald, 2010 ; IPCC, 2014 ). These brief episodes bring plants beyond their capacity of adaptation; decreasing crop and tree yield ( Ciais et al., 2005 ; Zinta et al., 2014 ).

Here we will not discuss plants capacity of adaptation to novel environmental conditions when considering large scales and long-term periods. Ecosystems are being affected by climate change at all levels (terrestrial, freshwater, and marine), and it was already reported that species are under evolutionary adaptation to human-caused climate change (for a review see Scheffers et al., 2016 ). Migration and plasticity are two biological mechanisms to cope with these changes. Data indicate that each population of a species has limited tolerance to sharp climate variations, and they could migrate to find more favorable environments. Habitat fragmentation limits plant movement, being other big threat for adaptation ( Stockwell et al., 2003 ; Leimu et al., 2010 ). Despite the fact that individual plants are immobile, plant populations move when seeds are dispersed, resulting in differences in the general distribution of the species ( Corlett and Westcott, 2013 ). In this sense, anthropogenic activities also contribute to seed dispersal.

Plasticity is a characteristic related to phenology and phenotype. Phenology is the timing of phases occurrence in the life cycle, and phenotypic plasticity is the range of phenotypes that a single genotype may express depending on its environment ( Nicotra et al., 2010 ). Plasticity is adaptive when the phenotype changes occur in a direction favored by selection in the new environment.

Climate Change and ROS

Reactive Oxygen Species (ROS) are continuously generated by plants under normal conditions. However, they are increased in response to different abiotic stresses. One of the most important effects of climate change-related stresses at the molecular level is the increase of ROS inside the cells ( Farnese et al., 2016 ). Among ROS, the most studied are superoxide anion ( O 2 •– ), H 2 O 2 and the hydroxyl radical (⋅OH - ).

Reactive Oxygen Species cause damage to proteins, lipids and DNA, affecting cell integrity, morphology, physiology, and, consequently, the growth of plants ( Frohnmeyer and Staiger, 2003 ). The main sources of ROS in stress conditions are: augmented photorespiration, NADPH oxidase (NOX) activity, β-oxidation of fatty acids and disorders in the electron transport chains of mitochondrias and chloroplasts ( Apel and Hirt, 2004 ; AbdElgawad et al., 2015 ). Hence, higher plants have evolved in the presence of ROS and have acquired pathways to protect themselves from its toxicity. Plant antioxidant system (AS) includes the activity of ROS detoxifying enzymes [e.g., superoxide dismutase (SOD), ascorbate peroxidase (APX), catalase (CAT), glutathione peroxidase (GPX), and peroxiredoxin (PRX)], as well as antioxidant molecules such as ascorbic acid (ASC) and glutathione (GSH) that are present in almost all subcellular compartments (reviewed by Choudhury et al., 2017 ).

In this context, plants have also developed a tight interaction between ROS and NO as a mechanism to reduce the deleterious consequences of these ROS-induced oxidative injuries. NO orchestrates a wide range of mechanisms leading to the preservation of redox homeostasis in plants. Consequently, NO at low concentration is considered a broad-spectrum anti-stress molecule ( Lamattina et al., 2003 ; Tossi et al., 2009 ; Correa-Aragunde et al., 2015 ). Figure 1 shows the relationship among the different GHG and their impact on plants.

www.frontiersin.org

FIGURE 1. Simplified scheme showing greenhouse gasses (GHG) and their effects on plants. GHG (H 2 O vapor, clouds, CO 2 , CH 4 , N 2 O, and NO) have both natural and anthropogenic origin, contributing to greenhouse effect. Short-term effects of GHG increase is mainly CO 2 rise, that activates photosynthesis (PS) and inhibits stomatal opening (SO). Long-term effects of GHG increase are extreme climate changes such as floods, droughts, heat. All of them induce the generation of reactive oxygen species (ROS) and oxidative stress in plants. Nitric oxide (NO) could alleviate oxidative stress by scavenging ROS and/or regulating the antioxidant system (AS). GHG and volatile organic compounds (VOC) react in presence of sunlight (E#) to give tropospheric O 3 . Although tropospheric O 3 is prejudicial for life, stratospheric O 3 is beneficial, because filters harmful UV-B radiation. The size of arrows are representative of the GHG concentration.

CO 2 and NO Contribute to Regulate Redox Homeostasis in Plants

Co 2 increasing: advantages and disadvantages.

Increased CO 2 was suggested to have a “fertilization” effect, because crops would increase their photosynthesis and stomatal conductance in response to elevated CO 2 . This belief was supported by studies performed in greenhouses, laboratory controlled-environment chambers, and transparent field chambers, where emitted CO 2 may be held back and readily controlled ( Drake et al., 1997 ; Markelz et al., 2014 ). However, more realistic results, obtained by Free-Air Concentration Enrichment (FACE) technology, suggest that the fertilization response due to CO 2 increase is probably dependent on genetic and environmental factors, and the duration of the study ( Smith and Dukes, 2013 ). An extensive review of the literature in this field made by Xu et al. (2015) concluded that augmented CO 2 normally increases photosynthesis in C3 species such as rice, soybean and wheat. On the other hand, they pointed out that a negative feedback of photosynthesis could take place in augmented CO 2 , as a result of overload of chemical and reactive generated substrates, leading to an imbalance in the sink:source carbon ratio. Moreover, the energetic cost of carbohydrate exportation increases in elevated CO 2 level.

The most important photosynthetic enzyme is the ribulose-1,5-bisphosphate carboxylase-oxygenase (RuBisCO). Rubisco is located in mesophyll cells of C3 plants, in direct contact with the intercellular air space linked to the atmosphere by epidermal stomatal pores. Photosynthesis increases at high CO 2 , because Rubisco is not CO 2 saturated and CO 2 inhibits the oxygenation reactions and photorespiration ( Long et al., 2006 ). However, long-term high concentration of CO 2 may down regulate Rubisco activity because ribulose-1,5-bisphosphate is not regenerated. Hexokinase (HXK), a sensor of extreme photosynthate, may participate in the down regulation of Rubisco concentration ( Xu et al., 2015 ). Moreover, severe abiotic stresses, such as temperature and drought, may restrain Rubisco carboxylation and foster oxygenation ( Xu et al., 2015 ).

In C4 crops, such as maize and sorghum, the elevated concentration of CO 2 inside the bundle sheath cells could prevent a large increase of Rubisco activity at higher atmospheric CO 2 and, thereby, photosynthetic activity is not augmented. However, at high CO 2 levels, the water status of C4 plants under drought conditions is improved, increasing photosynthesis and biomass accumulation ( Long et al., 2006 ; Mittler and Blumwald, 2010 ). That envisages potential advantages for the C4 species in future climatic change scenarios, particularly in arid and semiarid areas.

In addition, high CO 2 has the benefit of reducing stomatal conductance, decreasing 10% evapotranspiration in both C3 and C4 plants. Simultaneously, the cooling decreased resulting from reduced transpiration causes elevated canopy temperatures of around 0.7°C for most crops. Biomass and yield rise due to high CO 2 in all C3 plants, but not in C4 plants exception made when water is a restraint. Yields of C3 grain crops jump around 19% on average at high CO 2 ( Kimball, 2016 ).

Some reports analyze the contribution of CO 2 in the responses of plants to the combination of multiple stresses. For Arabidopsis thaliana , the combination of heat and drought induces photosynthesis inhibition of 62% under ambient CO 2 , but the drop in photosynthesis is just 40% at high CO 2 . Moreover, the protein oxidation increases significantly during a heat wave and drought, and this effect is repressed by increased CO 2 . Photorespiration is also reduced by high CO 2 ( Zinta et al., 2014 ).

Studying grasses ( Lolium perenne, Poa pratensis ) and legumes ( Medicago lupulina, Lotus corniculatus ) exposed to drought, high temperature and augmented CO 2 , AbdElgawad et al. (2015) demonstrated that drought suppresses plant growth, photosynthesis and stomatal conductance, and promotes in all species the synthesis of osmolytes and antioxidants. Instead, oxidative damage is more markedly observed in legumes than in grasses. In general, warming amplifies drought consequences. In contrast, augmented CO 2 diminishes stress impact. Reduction in photosynthesis and chlorophyll, as a result of drought and elevated temperature, were avoided by high CO 2 in the grasses. Noxious effects of oxidative stress, i.e., lipid peroxidation, are phased down in all species by augmented CO 2 . Normally, a reduced impact of oxidative stress is due to decreased photorespiration and diminished NOX activity. In legumes, a rise in levels of antioxidant molecules (flavonoids and tocopherols) contribute as well to the stress mitigation caused by augmented CO 2 . The authors draw the conclusion that these different responses point at an unequal future impact of climate change on the production of agricultural-scale legumes and grass crops.

Kumari et al. (2015) assessed the impact of various levels of CO 2 , ambient (382 ppm) and augmented (570 ppm), and O 3 , ambient (50 ppb) and augmented (70 ppb) on the potato physiological and biochemical responses ( Solanum tuberosum ). They observed that augmented CO 2 cut down O 3 uptake, enhanced carbon assimilation, and curbed oxidative stress. Elevated CO 2 also mitigated the noxious effect of high O 3 on photosynthesis.

Although some molecular mechanisms underpinning CO 2 actions are unknown, the results presented highlight the importance of CO 2 as a regulator that mitigates the potential climate change-induced deleterious consequences in plants. Recent reports suggest that some CO 2 -associated responses may be mediated by NO.

Du et al. (2016) determined that 800 μmol.mol -1 of CO 2 increased the NO concentration in Arabidopsis leaves, through a mechanism related to nitrate availability. Moreover, NO increase, as a consequence of high CO 2 levels, was reported as a general procedure to improve iron (Fe) nutrition in response to Fe deficiency in tomato roots ( Jin et al., 2009 ).

The gas exchange between the atmosphere and plants is mainly regulated by stomata. But structure and physiology of stomata are also influenced by gasses ( García-Mata and Lamattina, 2013 ). Elevated CO 2 regulate stomatal density and conductance. Moreover, there is increasing evidence that this response is modified by interaction of CO 2 with other environmental factors ( Xu et al., 2016 ; Yan et al., 2017 ). Wang et al. (2015) reported that 800 μmol.mol -1 of CO 2 increases the NO concentration in A. thaliana guard cells, inducing stomatal closure. Both NR and NO synthase (NOS)-like activities are necessary for CO 2 -induced NO accumulation. Comprehensive pharmacological and genetic results obtained in Arabidopsis by Chater et al. (2015) , show that when CO 2 concentration is around 700–1000 ppm, stomatal density and closure are reduced. They also illustrate that those elements necessary for this process are: activation of both ABA biosynthesis genes and the PYR/RCAR ABA receptor, and ROS increase. However, Shi et al. (2015) provide genetic and pharmacological evidence that high CO 2 concentration induces stomatal closure by an ABA-independent mechanism in tomato. They show that 800 μmol.mol -1 of CO 2 increase the expression of the protein kinase OPEN STOMATA 1 (OST1), NOX, and nitrate reductase (NR) genes. They also show that the sequential production of NOX-dependent H 2 O 2 and NR-produced NO are mainly dependent of OST1, and are involved in the CO 2 -induced stomatal closure.

In ABA-dependent mechanisms, ABA is increased by CO 2. The binding of ABA to its receptor (PYR/RCAR) inactivates PP2C, activating OST1. In ABA-independent mechanism, OST1 will be transcriptionally induced by CO 2 . Once activated, OST1 along with Ca 2 + , activates NOX, increasing ROS ( Kim et al., 2010 ). The rise of guard cells ROS enhances NO, cytosolic free Ca 2 + , and pH ( Song et al., 2014 ; Xie et al., 2014 ). ROS and NO release Ca 2 + from internal reservoirs, or influx external Ca 2 + through plasma membrane Ca 2 + in channels. Cytosolic free Ca 2 + inactivate inward K + channels (K + in ) to prevent K + uptake and activate outward K + channels (K + out ) and Cl - (anion) channels (Cl - ) at the plasma membrane ( Blatt, 2000 ; García-Mata et al., 2003 ). Ca 2 + also activates slow anion channel homolog 3 (SLAH3), slow anion channel-associated 1 (SLAC1) and aluminum activated malate transporters (ALMT) ( Roelfsema et al., 2012 ). The consequence of the regulation of cation/anion channels is the net efflux of K + /Cl - /malate and influx of Ca 2 + , making guard cells lose turgor by water outlet, causing stomatal closure.

All together, the results discussed here suggest that CO 2 -induced NO increase is a common plant physiological response to oxidative stresses. Figure 2 shows the importance of CO 2 and NO in these processes.

www.frontiersin.org

FIGURE 2. Interplay between CO 2 and NO in plant redox physiology: CO 2 enters to the leaves by stomata. Once in mesophyll cells, CO 2 increase photosynthesis (PS) through the CO 2 -unsaturated Rubisco activity. When plants are in stress environments, ROS could be augmented by Rubisco-induced photorespiration and NADPH oxidase (NOX) activities. NOX- induced O 2 •– , in the apoplast is immediately transformed to H 2 O 2 by the superoxide dismutase (SOD). Plasma membrane is permeable to H 2 O 2 . CO 2 moderates oxidative stress in mesophyll cells by inhibiting both Rubisco photorespiration (PR) and NOX activities. Besides, NO is induced by CO 2 and ROS, alleviating the consequences of oxidative stress by scavenging ROS and activating or inhibiting the antioxidant system (AS). In guard cells, CO 2 increases the expression and activity of OPEN STOMATA 1 (OST1), in both ABA-dependent and independent mechanisms. OST1 activates NOX, producing ROS and consequently NO increase by nitrate reductase (NR), and NOS-like activities. NO prevents ROS increase by direct scavenging, and inhibiting NOX. NO-dependent Ca 2 + regulated ion channels induces stomatal closure, modulating O 3 and CO 2 uptake, decreasing evapotranspiration, and rising leaf temperature.

Abiotic Stress, ROS Generation, and Redox Balance: The Key Role of NO

Reactive oxygen species are generated in apoplast, plasma membrane, chloroplasts, mitochondria, and peroxisomes ( Farnese et al., 2016 ). It was proposed that each stress produces its own “ROS signature” ( Choudhury et al., 2017 ). For instance, drought may reduce the activity of Rubisco, decreasing CO 2 fixation and NADP+ regeneration by the Calvin cycle. As a consequence, chloroplast electron transport is altered, generating ROS by electron leakage to O 2 ( Carvalho, 2008 ). In drought stress, ROS increase is produced by NOX activity ( Farnese et al., 2016 ). In flooding, ROS generation is an ethylene-promoted process that involves calcium (Ca 2+ ) flux, and NOX activity ( Voesenek and Bailey-Serres, 2015 ).

In heat stress, a NOX-dependent transient ROS rise is an early event ( Königshofer et al., 2008 ). Then, endogenous ROS are sensed through histidine kinases, and an Arabidopsis heat stress factor (HsfA4a) appears to sense exogenous ROS. As a result, the MAPK signal pathway is activated ( Qu et al., 2013 ). Moreover, functional decrease in photosynthetic light reaction induces ROS concentration by high electron leakage from the thylakoid membrane ( Hasanuzzaman et al., 2013 ). In this process, O 2 is the acceptor, generating O 2 •– .

Thus, individual stresses or their different combinations may produce particular “ROS signatures.” Besides their deleterious effects, ROS are recognized as a signal in the plant reaction to biotic and abiotic stressors. ROS may induce programed cell death (PCD) to avoid pathogen spread ( Mur et al., 2008 ), trigger a systemic defense response signal ( Dubiella et al., 2013 ), or avoid the chloroplast antenna overloading by electrons divert ( Choudhury et al., 2017 ).

Whatever the origin and function, ROS concentration must be adequately regulated to avoid excessive concentration and consequent cellular damages. Depending on NO and ROS concentrations, NO has the dual capacity to activate or inhibit the ROS production, and is a key molecule for keeping cellular redox homeostasis under control ( Beligni and Lamattina, 1999a ; Correa-Aragunde et al., 2015 ). NO has a direct ROS-scavenging activity because it holds an unpaired electron, reaching elevated reactivity with O 2 , O 2 •– , and redox active metals. NO can mitigate OH formation by scavenging either Fe or O 2 •– ( Lamattina et al., 2003 ). However, NO reacting with ROS (mainly O 2 •– ) may generate reactive nitrogen species (RNS). An excess of RNS originates a nitrosative stress ( Corpas et al., 2011 ). To avoid the toxicity of nitrosative stress, NO is stored as GSNO in the cell.

GSH as a Redox Buffer. GSNO as NO Reservoir. SNO and S-Nitrosylation

Glutathione (GSH) is a small peptide with the sequence γ-l-glutamyl-l-cysteinyl-glycine that has a cell redox homeostatic impact in most plant tissues. It is a soluble small thiol considered a non-enzymatic antioxidant. It exists in the reduced (GSH) or oxidized state (GSSG), in which two GSH molecules are joined by a disulfide bond ( Rouhier et al., 2008 ). GSH alleviates oxidative damages in plants generated by abiotic stresses, including salinity, drought, higher, low temperature, and heavy metals. GSH is precursor of phytochelatins, polymers that chelate toxic metals and transport them to the vacuole ( Grill et al., 1989 ). Studies shown that GSH contributes to tolerate nickel, cadmium, zinc, mercury, aluminum and arsenate heavy metals in plants ( Asgher et al., 2017 ). Moreover, GSH has a role in the detoxification of ROS both directly, interacting with them, or indirectly, participating of enzymatic pathways. GSH is involved in glutathionylation, a posttranslational modification that causes a mixed disulfide bond between a Cys residue and GSH.

GSH can be oxidized to GSSG by H 2 O 2 and can react with NO to form the nitrosoglutathione (GSNO) derivative. GSNO is an intracellular NO reservoir. It is also a vehicle of NO throughout the cell and organs, spreading NO biological function. GSNO is the largest low-molecular-mass S-nitrosothiol (SNO) in plant cells ( Corpas et al., 2013 ). GSNO metabolism and its reaction with other molecules involve S-nitrosylation and S-transnitrosation which consist of the binding of a NO molecule to a cysteine residue in proteins. Thioredoxin produces protein denitrosylation ( Correa-Aragunde et al., 2013 ). GSNO could be decomposed by the GSNO reductase (GSNOR) to GSSG which, in turn, is reduced to GSH by glutathione reductase (GR).

Glutathione also participates in the GSH/ASC cycle, a series of enzymatic reactions that degrade H 2 O 2 . APX degrades H 2 O 2 using ASC, the other major antioxidant in plants, as cofactor. The oxidized ASC is reduced by monodehydroascorbate reductase (MDHAR) in an NAD(P)H-dependent manner and by dehydroascorbate reductase (DHAR) employing GSH as electron donor. The resulting GSSG is reduced in turn to GSH by GR ( Foyer and Noctor, 2011 ).

Different Effects of NO in the Regulation of Antioxidant Enzymes

The application of NO donors alleviates oxidative stress in plants challenged to abiotic and/or biotic stresses ( Laxalt et al., 1997 ; Beligni and Lamattina, 1999b , 2002 ; Shi et al., 2007 ; Xue et al., 2007 ; Leitner et al., 2009 ).

Besides the direct ROS-scavenging activity of NO, its beneficial effect is exerted by the regulation of the antioxidant enzymes activity that controls toxic levels of ROS and RNS ( Uchida et al., 2002 ; Shi et al., 2005 ; Song et al., 2006 ; Romero-Puertas et al., 2007 ; Bai et al., 2011 ). NO can modulate cell redox balance in plants through the regulation of gene expression, posttranslational modification or by its binding to the heme prosthetic group of some antioxidant enzymes.

SOD catalyzes the dismutation of stress-generated O 2 •– in one of two less harmful species: either molecular oxygen (O 2 ) or hydrogen peroxide (H 2 O 2 ). APX and CAT are the most important enzymes degrading H 2 O 2 in plants. They transform H 2 O 2 to H 2 O and O 2 . APX isoforms are primarily found in the cytosol and chloroplasts, while the CAT isoforms are found in peroxisomes. APX has strong affinity for H 2 O 2 and uses ASC as an electron donor. In contrast, CAT removes H 2 O 2 generated in the peroxisomal respiratory pathway without the need to reduce power. Even though CAT affinity for H 2 O 2 is low, its elevated rate of reaction offers an effective way to detoxify H 2 O 2 inside the cell. PRX may reduce both hydroperoxide and peroxynitrite.

Many reports on different plant species demonstrate that NO induces the transcription and activity of antioxidative enzymes in response to oxidative stress. The tolerance to drought and salt-induced oxidative stress in tobacco is related to the ABA-triggered production of H 2 O 2 and NO. In turn, they induce transcripts and activities of SOD, CAT, APX, and GR ( Zhang et al., 2009 ). UV-B-produced oxidative stress in Glycine max was alleviated by NO donors, which induced transcription and activities of SOD, CAT, and APX ( Santa-Cruz et al., 2014 ). Furthermore, in bean leaves, SOD, CAT, and APX activities are increased by NO donors, and protected from the oxidative stress generated by UV-B irradiation ( Shi et al., 2005 ). Drought tolerance in bermudagrass is improved by ABA-dependent SOD and CAT activities. This effect is regulated by H 2 O 2 and NO, NO acting downstream H 2 O 2 ( Lu et al., 2009 ).

Several antioxidant enzymes have been identified as target of S-nitrosylation, resulting in a change of their biological activity ( Romero-Puertas et al., 2008 ; Bai et al., 2011 ; Fares et al., 2011 ). For instance, NO reinforces recalcitrant seed desiccation tolerance in Antiaris toxicaria by activating the ascorbate-glutathione cycle through S-nitrosylation to control H 2 O 2 accumulation. Desiccation treatment reduced the level of S-nitrosylated APX, GR, and DHAR proteins. Instead, NO gas exposure activated them by S-nitrosylation ( Bai et al., 2011 ). Furthermore, APX was S-nitrosylated at Cys32 during saline stress and biotic stress, enhancing its enzymatic activity ( Begara-Morales et al., 2014 ; Yang et al., 2015 ). In addition, auxin-induced denitrosylation of cytosolic APX provoked inhibition of its activity, followed by an increase of H 2 O 2 concentration and the consequent lateral root formation in Arabidopsis ( Correa-Aragunde et al., 2013 ). Moreover, an inhibitory impact of S-nitrosylation on APX activity was also reported during programmed cell death in Arabidopsis ( de Pinto et al., 2013 ). CAT was identified to be S-nitrosylated in a proteomic study of isolated peroxisomes ( Ortega-Galisteo et al., 2012 ). A decrease of S-nitrosylated CAT under Cd treatment was reported. In addition, in vitro experiments demonstrated a reversible inhibitory effect of APX and CAT activities by NO binding to the Fe of the heme cofactor ( Brown, 1995 ; Clark et al., 2000 ). In addition, NOXs have been involved in plant defense, development, hormone biosynthesis and signaling ( Marino et al., 2012 ). Whereas S-nitrosylation did not affect SOD activities, nitration inhibited Mn-SOD1, Fe-SOD3, and CuZn-SOD3 activity to different degrees ( Holzmeister et al., 2015 ). SOD isoforms could also regulate endogenous NO availability by competing for the common substrate, O 2 •– , and it was demonstrated that bovine SOD may release NO from GSNO ( Singh et al., 1999 ). When GSNO is decomposed by GSNOR, it produces GSSG. GSNOR is also regulated by NO. Frungillo et al. (2014) demonstrated that NO-derived from nitrate assimilation in Arabidopsis inhibited GSNOR1 by S-nitrosylation, preventing GSNO degradation. They proposed that (S)NO controls its own generation and scavenging by modulating nitrate assimilation and GSNOR1 activity. It was also shown that chilling treatment in poplar increased S-nitrosylation of NR, along with a significant decrease of its activity ( Cheng et al., 2015 ).

The dual activity of Prx, suggests a role for this enzyme both in ROS and RNS regulation. S-nitrosylation of Arabidopsis PrxIIE inhibits its peroxynitrite activity, increasing peroxynitrite-mediated tyrosine nitration ( Romero-Puertas et al., 2007 ). Pea mitochondrial PrxIIF was S-nitrosylated under salt stress, and its peroxidase activity was reduced by 5 mM GSNO ( Camejo et al., 2013 ).

An interesting study demonstrated that NO controls hypersensitive response (HR) through S-nitrosylation of NOX, inhibiting ROS synthesis. This triggers a feedback loop limiting HR ( Yun et al., 2011 ).

Other proteins related to abiotic stress response are regulated by S-nitrosylation (For a review see Fancy et al., 2017 ).

Figure 3 is a simplified diagram that illustrates the main oxidative and nitrosative effects that modulate the activities of key cell components, thus maintaining cell redox balance. Note the feedback and positive-negative regulatory processes occurring in the main pathways. They involve posttranslational modifications that activate and inhibit the components involved in cell antioxidant system.

www.frontiersin.org

FIGURE 3. Molecules and mechanisms involved in NO-mediated redox balance. H 2 O 2 is generated mainly by NOX and SOD as a response to (a)biotic stress. APX and CAT are the main H 2 O 2 -degrading enzymes. NO is increased by H 2 O 2 through the induction of NR/NOS-like activities, and may scavenge ROS or induce both the transcription and activity of SOD, CAT, and APX. In parallel, NO is combined with GSH to form nitrosoglutathione GSNO. GSNO regulates many enzymatic activities by the posttranslational modification of cysteine residues through S-Nitrosylation. NOX and CAT activities are inhibited by S-nitrosylation, whereas APX is either activated or inhibited by S-nitrosylation. NO also inhibits APX by binding to heme group. GSNO is degraded by GSNOR, which could be inhibited by H 2 O 2 and S-nitrosylation.NR could be inhibited by S-nitrosylation. GR reduces GSSG to GSH, and it is activated by S-nitrosylation. Ascorbate (ASC) is a cofactor of APX. Reduced ASC is generated by MDHAR and DHAR, using GSH as electron donor. Both enzymes are inhibited by S-nitrosylation. Reactive Nitrogen Species (RNS) may be originated by NO and O 2 •– reaction. SOD regulate RNS dismutating O 2 •– . Peroxiredoxins (Prx) reduce both ROS AND RNS. RNS are degraded by PrxIIe, and H 2 O 2 by PrxIIF. Both enzymes are inhibited by S-nitrosylation. Red lines: H 2 O 2 -regulated reactions. Purple lines: NO-regulated reactions. Green lines: GSNO-regulated reactions.

Conclusions and Perspectives

The accelerating rate of climate change, together with habitat fragmentation caused by human activity, are part of the selective pressures building a new Earth’s landscape.

Climate change is a multidimensional and simultaneous variation in duration, frequency and intensity of parameters like temperature and precipitation, altering the seasons and life on the Earth. In this scenario, plant species with increased adaptive plasticity will be better equipped to tolerate changes in the frequency of extreme weather events. GHG are one of the forces driving climate change. However, CO 2 and NO may contribute to maintaining the cell redox homeostasis, regulating the amount of ROS, GSH, GSNO, and SNO.

In this manuscript, we summarize the available evidence supporting the presence of broad spectrum anti-stress molecules, as NO in plants, for coping with unprecedented changes in environmental conditions. Future research should focus in better understanding the influence of GHG on plant physiology.

Author Contributions

RC conceived the project and wrote the manuscript. MN drew figures and collaborated in writing the manuscript. NC-A and LL supervised and complemented the drafting. All the persons entitled to authorship have been named and have approved the final version of the submitted manuscript.

This work was supported by grants from the Consejo Nacional de Investigaciones Cientificas y Tecnicas, the Agencia Nacional de Promoción Científica y Tecnológica, and the Universidad Nacional de Mar del Plata, Argentina. NC-A, LL, and RC are permanent members of the Scientific Research career of CONICET. MN is doctoral fellow of the ANPCYT.

Conflict of Interest Statement

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

The reviewer MCR-P and handling Editor declared their shared affiliation.

Acknowledgments

We thank ANPCYT for MN fellowship. We also thank Marta Terrazo for helping with the language revision of the manuscript.

  • ^ https://ozonewatch.gsfc.nasa.gov/facts/ozone.html

AbdElgawad, H., Farfan-Vignolo, E. R., de Vos, D., and Asard, H. (2015). Elevated CO2 mitigates drought and temperature-induced oxidative stress differently in grasses and legumes. Plant Sci. 231, 1–10. doi: 10.1016/j.plantsci.2014.11.001

PubMed Abstract | CrossRef Full Text | Google Scholar

Apel, K., and Hirt, H. (2004). Reactive oxygen species: metabolism, oxidative stress, and signal transduction. Annu. Rev. Plant Biol. 55, 373–399. doi: 10.1146/annurev.arplant.55.031903.141701

Asgher, M., Per, T. S., Anjum, S., Khan, M., Masood, A., Verna, S., et al. (2017). “Contribution of glutathione in heavy metal stress tolerance in plants,” in Reactive Oxygen Species and Antioxidant Systems in Plants: Role and Regulation under Abiotic Stress , eds M. Khan and N. Khan (Singapore: Springer), doi: 10.1007/978-981-10-5254-5_12

CrossRef Full Text | Google Scholar

Bai, X., Yang, L., Tian, M., Chen, J., Shi, J., Yang, Y., et al. (2011). Nitric oxide enhances desiccation tolerance of recalcitrant Antiaris toxicaria seeds via protein S-nitrosylation and carbonylation. PLoS One 6:e20714. doi: 10.1371/journal.pone.0020714

Begara-Morales, J. C., Sánchez-Calvo, B., Chaki, M., Valderrama, R., Mata-Pérez, C., López-Jaramillo, J., et al. (2014). Dual regulation of cytosolic ascorbate peroxidase (APX) by tyrosine nitration and S -nitrosylation. J. Exp. Bot. 65, 527–538. doi: 10.1093/jxb/ert396

Beligni, M. V., and Lamattina, L. (1999a). Is nitric oxide toxic or protective? Trends Plant Sci. 4, 299–300. doi: 10.1016/S1360-1385(99)01451-X

Beligni, M. V., and Lamattina, L. (1999b). Nitric oxide protects against cellular damage produced by methylviologen herbicides in potato plants. Nitric Oxide 3, 199–208. doi: 10.1006/niox.1999.0222

Beligni, M. V., and Lamattina, L. (2002). Nitric oxide interferes with plant photo-oxidative stress by detoxifying reactive oxygen species. Plant Cell Environ. 25, 737–748. doi: 10.1046/j.1365-3040.2002.00857.x

Blatt, M. R. (2000). Cellular signaling and volume control in stomatal movements in plants. Annu. Rev. Cell Dev. Biol. 16, 221–241. doi: 10.1146/annurev.cellbio.16.1.221

Brown, G. C. (1995). Reversible binding and inhibition of catalase by nitric oxide. Eur. J. Biochem. 232, 188–191. doi: 10.1111/j.1432-1033.1995.tb20798.x

Camejo, D., Romero-Puertas, M. D. C., Rodríguez-Serrano, M., Sandalio, L. M., Lázaro, J. J., Jiménez, A., et al. (2013). Salinity-induced changes in S-nitrosylation of pea mitochondrial proteins. J. Proteomics 79, 87–99. doi: 10.1016/j.jprot.2012.12.003

Carvalho, M. H. C. (2008). Drought stress and reactive oxygen species. Plant Signal. Behav. 3, 156–165. doi: 10.4161/psb.3.3.5536

Chamizo-Ampudia, A., Sanz-Luque, E., Llamas, A., Ocana-Calahorro, F., Mariscal, V., Carreras, A., et al. (2016). A dual system formed by the ARC and NR molybdoenzymes mediates nitrite-dependent NO production in Chlamydomonas . Plant Cell Environ. 39, 2097–2107. doi: 10.1111/pce.12739

Cheng, T., Chen, J., Ef, A. A., Wang, P., Wang, G., Hu, X., et al. (2015). Quantitative proteomics analysis reveals that S-nitrosoglutathione reductase (GSNOR) and nitric oxide signaling enhance poplar defense against chilling stress. Planta 242, 1361–1390. doi: 10.1007/s00425-015-2374-5

Choudhury, F. K., Rivero, R. M., Blumwald, E., and Mittler, R. (2017). Reactive oxygen species, abiotic stress and stress combination. Plant J. 90, 856–867. doi: 10.1111/tpj.13299

Ciais, P., Reichstein, M., Viovy, N., Granier, A., Ogée, J., Allard, V., et al. (2005). Europe-wide reduction in primary productivity caused by the heat and drought in 2003. Nature 437, 529–533. doi: 10.1038/nature03972

Clark, D., Durner, J., Navarre, D. A., and Klessig, D. F. (2000). Nitric oxide inhibition of tobacco catalase and ascorbate peroxidase. Mol. Plant. Microbe. Interact. 13, 1380–1384. doi: 10.1006/niox.1999.0222

Corlett, R. T., and Westcott, D. A. (2013). Will plant movements keep up with climate change? Trends Ecol. Evol. 28, 482–488. doi: 10.1016/j.tree.2013.04.003

Corpas, F. J., Alché, J. D., and Barroso, J. B. (2013). Current overview of S-nitrosoglutathione (GSNO) in higher plants. Front. Plant Sci. 4:126. doi: 10.3389/fpls.2013.00126

Corpas, F. J., Leterrier, M., Valderrama, R., Airaki, M., Chaki, M., Palma, J. M., et al. (2011). Nitric oxide imbalance provokes a nitrosative response in plants under abiotic stress. Plant Sci. 181, 604–611. doi: 10.1016/j.plantsci.2011.04.005

Correa-Aragunde, N., Foresi, N., Delledonne, M., and Lamattina, L. (2013). Auxin induces redox regulation of ascorbate peroxidase 1 activity by S-nitrosylation/denitrosylation balance resulting in changes of root growth pattern in Arabidopsis . J. Exp. Bot. 64, 3339–3349. doi: 10.1093/jxb/ert172

Correa-Aragunde, N., Foresi, N., and Lamattina, L. (2015). Nitric oxide is a ubiquitous signal for maintaining redox balance in plant cells: regulation of ascorbate peroxidase as a case study. J. Exp. Bot. 66, 2913–2921. doi: 10.1093/jxb/erv073

Chater, C., Peng, K., Movahedi, M., Dunn, J. A., Walker, H. J., Liang, Y. K., et al. (2015). Elevated CO 2 -induced responses in stomata require ABA and ABA signaling. Curr. Biol. 25, 2709–2716. doi: 10.1016/j.cub.2015.09.013

de Pinto, M. C., Locato, V., Sgobba, A., Romero-Puertas, M. D. C., Gadaleta, C., Delledonne, M., et al. (2013). S-Nitrosylation of ascorbate peroxidase is part of programmed cell death signaling in tobacco bright yellow-2 cells. Plant Physiol. 163, 1766–1775. doi: 10.1093/jxb/ert172

Di Dato, V., Musacchia, F., Petrosino, G., Patil, S., Montresor, M., Sanges, R., et al. (2015). Transcriptome sequencing of three Pseudo-nitzschia species reveals comparable gene sets and the presence of nitric oxide synthase genes in diatoms. Sci. Rep. 5:12329. doi: 10.1038/srep12329

Drake, B. G., Gonzalez-Meler, M. A., and Long, S. P. (1997). More efficient plants: a consequence of rising atmospheric CO 2 ? Annu. Rev. Plant Physiol. Plant Mol. Biol. 48, 609–639. doi: 10.1146/annurev.arplant.48.1.609

Du, S., Zhang, R., Zhang, P., Liu, H., Yan, M., Chen, N., et al. (2016). Elevated CO 2 -induced production of nitric oxide (NO) by NO synthase differentially affects nitrate reductase activity in Arabidopsis plants under different nitrate supplies. J. Exp. Bot. 67, 893–904. doi: 10.1093/jxb/erv506

Dubiella, U., Seybold, H., Durian, G., Komander, E., Lassig, R., Witte, C. P., et al. (2013). Calcium-dependent protein kinase/NADPH oxidase activation circuit is required for rapid defense signal propagation. Proc. Natl. Acad. Sci. U.S.A. 110, 8744–8749. doi: 10.1073/pnas.1221294110

Fancy, N. N., Bahlmann, A. K., and Loake, G. J. (2017). Nitric oxide function in plant abiotic stress. Plant Cell Environ. 40, 462–472. doi: 10.1111/pce.12707

Fares, A., Rossignol, M., and Peltier, J. (2011). Proteomics investigation of endogenous S-nitrosylation in Arabidopsis . Biochem. Biophys. Res. Commun. 416, 331–336. doi: 10.1038/srep12329

Farnese, F. S., Menezes-Silva, P. E., Gusman, G. S., and Oliveira, J. A. (2016). When bad guys become good ones: the key role of reactive oxygen species and Nitric Oxide in the plant responses to abiotic stress. Front. Plant Sci. 7:471. doi: 10.3389/fpls.2016.00471

Foresi, N., Correa-Aragunde, N., Parisi, G., Calo, G., Salerno, G., and Lamattina, L. (2010). Characterization of a nitric oxide synthase from the plant kingdom: NO generation from the green alga Ostreococcus tauri is light irradiance and growth phase dependent. Plant Cell 22, 3816–3830. doi: 10.1105/tpc.109.073510

Foyer, C. H., and Noctor, G. (2011). Ascorbate and glutathione: the heart of the redox hub. Plant Physiol. 155, 2–18. doi: 10.1038/srep12329

Frohlich, A., and Durner, J. (2011). The hunt for plant nitric oxide synthase (NOS): Is one really needed? Plant Sci. 181, 401–404. doi: 10.1016/j.plantsci.2011.07.014

Frohnmeyer, H., and Staiger, D. (2003). Ultraviolet-B-radiation-mediated responses in plants. Balancing damage and protection. Plant Physiol. 133, 1420–1428. doi: 10.1104/pp.103.030049

Frungillo, L., Skelly, M. J., Loake, G. J., Spoel, S. H., and Salgado, I. (2014). S-nitrosothiols regulate nitric oxide production and storage in plants through the nitrogen assimilation pathway. Nat. Commun. 5:5401. doi: 10.1038/ncomms6401

García-Mata, C., Gay, R., Sokolovski, S., Hills, A., Lamattina, L., and Blatt, M. R. (2003). Nitric oxide regulates K + and Cl-channels in guard cells through a subset of abscisic acid-evoked signaling pathways. Proc. Natl. Acad. Sci. U.S.A. 100, 11116–11121. doi: 10.1073/pnas.1434381100

García-Mata, C., and Lamattina, L. (2013). Gasotransmitters are emerging as new guard cell signaling molecules and regulators of leaf gas exchange. Plant Sci. 201-202, 66–73. doi: 10.3389/fpls.2016.00277

Gregg, J., Jones, C., and Dawson, T. (2003). Urbanization effects on tree growth in the vicinity of New York City. Nature 424, 183–187. doi: 10.1038/nature01728

Grill, E., Löffler, S., Winnacker, E. L., and Zenk, M. H. (1989). Phytochelatins, the heavy-metal-binding peptides of plants, are synthesized from glutathione by a specific gamma-glutamylcysteine dipeptidyl transpeptidase (phytochelatin synthase). Proc. Natl. Acad. Sci. U.S.A. 86, 6838–6842. doi: 10.1038/srep12329

Gupta, K. J., and Igamberdiev, A. U. (2016). Reactive nitrogen species in mitochondria and their implications in plant energy status and hypoxic stress tolerance. Front. Plant Sci. 7:369. doi: 10.3389/fpls.2016.00369

Hall, S., Huber, D., and Grimm, N. (2008). Soil N 2 O and NO emissions from an arid, urban ecosystem. J. Geophys. Res. 113:G01016. doi: 10.1029/2007JG000523

Hasanuzzaman, M., Nahar, K., Alam, M. M., Roychowdhury, R., and Fujita, M. (2013). Physiological, biochemical, and molecular mechanisms of heat stress tolerance in plants. Int. J. Mol. Sci. 14, 9643–9684. doi: 10.3390/ijms14059643

Held, I., and Souden, B. (2000). Water vapor feedback and global warming. Annu. Rev. Energy Environ. 25, 441–475. doi: 10.1146/annurev.energy.25.1.441

Holzmeister, C., Gaupels, F., Geerlof, A., Sarioglu, H., Sattler, M., Durner, J., et al. (2015). Differential inhibition of Arabidopsis superoxide dismutases by peroxynitrite-mediated tyrosine nitration. J. Exp. Bot. 66, 989–999. doi: 10.1093/jxb/eru458

IPCC (2014). Climate Change 2014: Mitigation of Climate Change. Contribution of Working Group III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change , eds O. R. Edenhofer, Y. Pichs-Madruga, E. Sokona, S. Farahani, K. Kadner, A. Seyboth, et al. Cambridge: Cambridge University Press.

Google Scholar

Jeandroz, S., Wipf, D., Stuehr, D. J., Lamattina, L., Melkonian, M., Tian, Z., et al. (2016). Occurrence, structure, and evolution of nitric oxide synthase-like proteins in the plant kingdom. Sci. Signal. 9:re2. doi: 10.1126/scisignal.aad4403

Jhun, I., Coull, B. A., Zanobetti, A., and Koutrakis, P. (2015). The impact of nitrogen oxides concentration decreases on ozone trends in the USA. Air Qual. Atmos. Health 8, 283–292. doi: 10.1088/1748-9326/10/8/084009

Jin, C. W., Du, S. T., Chen, W. W., Li, G. X., Zhang, Y. S., and Zheng, S. J. (2009). Elevated carbon dioxide improves plants iron nutrition through enhancing the iron-deficiency-induced responses under iron-limited conditions in tomato. Plant Physiol. 150, 272–280. doi: 10.1104/pp.109.136721

Karl, T. R., and Trenberth, K. E. (2003). Modern global climate change. Science 302, 1719–1723. doi: 10.1126/science.1090228

Kiehl, J., and Trenberth, K. (1997). Earth’s annual global mean energy budget. Bull. Amer. Meteor. Soc. 78, 197–208.

Kim, T., Böhmer, M., Hu, H., Nishimura, N., and Schroeder, J. I. (2010). Guard cell signal transduction network: advances in understanding abscisic acid, CO 2 , and Ca 2+ signaling. Annu. Rev. Plant Biol. 61, 561–591. doi: 10.1146/annurev-arplant-042809-112226

Kimball, B. (2016). Crop responses to elevated CO 2 and interactions with H 2 O, N, and temperature. Curr. Opin. Plant Biol. 31, 36–43. doi: 10.1016/j.pbi.2016.03.006

Königshofer, H., Tromballa, H. W., and Löppert, H. G. (2008). Early events in signaling high-temperature stress in tobacco BY2 cells involve alterations in membrane fluidity and enhanced hydrogen peroxide production. Plant Cell Environ. 31, 1771–1780. doi: 10.1111/j.1365-3040.2008.01880.x

Kumar, A., Castellano, I., Patti, F. P., Palumbo, A., and Buia, M. C. (2015). Nitric oxide in marine photosynthetic organisms. Nitric Oxide 47, 34–39. doi: 10.1038/srep12329

Kumari, S., Agrawal, M., and Singh, A. (2015). Effects of ambient and elevated CO 2 and ozone on physiological characteristics, antioxidative defense system and metabolites of potato in relation to ozone flux. Environ. Exp. Bot. 109, 276–287. doi: 10.1016/j.envexpbot.2014.06.015

Lamattina, L., García-Mata, C., Graziano, M., and Pagnussat, G. (2003). Nitric oxide: the versatility of an extensive signal molecule. Annu. Rev. Plant Biol. 54, 109–136. doi: 10.1146/annurev.arplant.54.031902.134752

Laxalt, A. M., Beligni, M. V., and Lamattina, L. (1997). Nitric oxide preserves the level of chlorophyll in potato leaves infected by Phytophthora infestans. Eur. J. Plant Pathol. 103, 643–651. doi: 10.1023/A:1008604410875

Le Treut, H., Somerville, R., Cubasch, U., Ding, Y., Mauritzen, C., Mokssit, A., et al. (2007). “Historical overview of climate change science,” in Climate change 2007: The physical science basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change , eds S. Solomon, D. Qin, M. Manning, Z. Chen, M. Marquis, K. B. Averyt, et al. (Cambridge: Cambridge University Press).

Leimu, R., Vergeer, P., Angeloni, F., and Ouborg, N. (2010). Habitat fragmentation, climate change, and inbreeding in plants. Ann. N. Y. Acad. Sci. 1195, 84–98. doi: 10.1111/j.1749-6632.2010.05450.x

Leitner, M., Vandelle, E., Gaupels, F., Bellin, D., and Delledonne, M. (2009). NO signals in the haze: nitric oxide signalling in plant defence. Curr. Opin. Plant Biol. 12, 451–458. doi: 10.1016/j.pbi.2009.05.012

Long, S., Ainsworth, E., Leakey, A., Nösberger, J., and Ort, D. (2006). Food for thought: lower-than-expected crop yield stimulation with rising CO 2 concentrations. Science 312, 1918–1921. doi: 10.1126/science.1114722

Long, S., Ainsworth, E., Rogers, A., and Ort, D. (2004). Rising atmospheric carbon dioxide: plants FACE the future. Annu. Rev. Plant Biol. 55, 591–628. doi: 10.1146/annurev.arplant.55.031903.141610

Lu, S., Su, W., Li, H., and Guo, Z. (2009). Abscisic acid improves drought tolerance of triploid bermudagrass and involves H 2 O 2 - and NO-induced antioxidant enzyme activities. Plant Physiol. Biochem. 7, 132–138. doi: 10.1016/j.plaphy.2008.10.006

Marino, D., Dunand, C., Puppo, A., and Pauly, N. (2012). A burst of plant NADPH oxidases. Trends Plant Sci. 17, 9–15. doi: 10.1016/j.tplants.2011.10.001

Markelz, R. J., Lai, L. X., Vosseler, L. N., and Leakey, A. D. (2014). Transcriptional reprogramming and stimulation of leaf respiration by elevated CO 2 concentration is diminished, but not eliminated, under limiting nitrogen supply. Plant Cell Environ. 37, 886–988. doi: 10.1111/pce.12205

McAdam, E. L., Brodribb, T. J., and McAdam, S. A. (2017). Does ozone increase ABA levels by non-enzymatic synthesis causing stomata to close? Plant Cell Environ. 40, 741–747. doi: 10.1111/pce.12893

Medinets, S., Skiba, U., Rennenberg, H., and Butterbach-Bahl, K. (2015). A review of soil NO transformation: associated processes and possible physiological significance on organisms. Soil Biol. Biochem. 80, 92–117. doi: 10.1016/j.soilbio.2014.09.025

Meyer, C., Lea, U. S., Provan, F., Kaiser, W. M., and Lillo, C. (2005). Is nitrate reductase a major player in the plant NO (nitric oxide) game? Photosynth. Res. 83, 181–189. doi: 10.1016/j.tplants.2011.10.001

Mittler, R., and Blumwald, E. (2010). Genetic engineering for modern agriculture: challenges and perspectives. Annu. Rev. Plant Biol. 61, 443–462. doi: 10.1146/annurev-arplant-042809-112116

Mur, L. A., Kenton, P., Lloyd, A. J., Ougham, H., and Prats, E. (2008). The hypersensitive response; the centenary is upon us but how much do we know? J. Exp. Bot. 59, 501–520. doi: 10.1093/jxb/erm239

Nicotra, A., Atkin, O., Bonser, S., Davidson, A., Finnegan, E., Mathesius, U., et al. (2010). Plant phenotypic plasticity in a changing climate. Trends Plant Sci. 15, 684–692. doi: 10.1016/j.tplants.2010.09.008

Ortega-Galisteo, A. P., Rodríguez-Serrano, M., Pazmiño, D. M., Gupta, D. K., Sandalio, L. M., and Romero-Puertas, M. C. (2012). S-Nitrosylated proteins in pea ( Pisum sativum L.) leaf peroxisomes: changes under abiotic stress. J. Exp. Bot. 63, 2089–2103. doi: 10.1016/j.tplants.2011.10.001

Pilegaard, K. (2013). Processes regulating nitric oxide emissions from soils. Philos. Trans. R. Soc. Lond. B Biol. Sci. 368:20130126. doi: 10.1098/rstb.2013.0126

Qu, A. L., Ding, Y. F., Jiang, Q., and Zhu, C. (2013). Molecular mechanisms of the plant heat stress response. Biochem. Biophys. Res. Commun. 432, 203–207. doi: 10.1016/j.bbrc.2013.01.104

Rockel, P., Strube, F., Rockel, A., Wildt, J., and Kaiser, W. M. (2002). Regulation of nitric oxide (NO) production by plant nitrate reductase in vivo and in vitro. J. Exp. Bot. 53, 103–110. doi: 10.1093/jexbot/53.366.103

Roelfsema, M. R., Hedrich, R., and Geiger, D. (2012). Anion channels: master switches of stress responses. Trends Plant Sci. 17, 221–229. doi: 10.1016/j.tplants.2012.01.009

Romero-Puertas, M. C., Campostrini, N., Mattè, A., Righetti, P. G., Perazzolli, M., Zolla, L., et al. (2008). Proteomic analysis of S-nitrosylated proteins in Arabidopsis thaliana undergoing hypersensitive response. Proteomics 8, 1459–1469. doi: 10.1002/pmic.200700536

Romero-Puertas, M. C., Laxa, M., Mattè, A., Zaninotto, F., Finkemeier, I., Jones, A. M. E., et al. (2007). S-nitrosylation of peroxiredoxin II E promotes peroxynitrite-mediated tyrosine nitration. Plant Cell 19, 4120–4130. doi: 10.1105/tpc.107.055061

Rouhier, N., Lemaire, S. D., and Jacquot, J.-P. (2008). The role of glutathione in photosynthetic organisms: emerging functions for glutaredoxins and glutathionylation. Annu. Rev. Plant Biol. 59, 143–166. doi: 10.1146/annurev.arplant.59.032607.092811

Santa-Cruz, D., Pacienza, N., Zilli, C., Tomaro, M., Balestrasse, K., and Yannarelli, G. (2014). Nitric oxide induces specific isoforms of antioxidant enzymes in soybean leaves subjected to enhanced ultraviolet-B radiation. J. Photochem. Photobiol. B 141, 202–209. doi: 10.1016/j.jphotobiol.2014.09.019

Scheffers, B. R., De Meester, L., Bridge, T. C., Hoffmann, A. A., Pandolfi, J. M., Corlett, R. T., et al. (2016). The broad footprint of climate change from genes to biomes to people. Science 354:aaf7671. doi: 10.1126/science.aaf7671

Schmidt, G. A., Ruedy, R. A., Miller, R. L., and Lacis, A. A. (2010). Attribution of the present-day total greenhouse effect. J. Geophys. Res. 115:D20106. doi: 10.1029/2010JD014287

Shi, K., Li, X., Zhang, H., Zhang, G., Liu, Y., Zhou, Y., et al. (2015). Guard cell hydrogen peroxide and nitric oxide mediate elevated CO 2 -induced stomatali movement in tomato. New Phytol. 208, 342–353. doi: 10.1111/nph.13621

Shi, Q., Ding, F., Wang, X., and Wei, M. (2007). Exogenous nitric oxide protect cucumber roots against oxidative stress induced by salt stress. Plant Physiol. Biochem. 45, 542–550. doi: 10.1016/j.plaphy.2007.05.005

Shi, S., Wang, G., Wang, Y., Zhang, L., and Zhang, L. (2005). Protective effect of nitric oxide against oxidative stress under ultraviolet-B radiation. Nitric Oxide 13, 1–9. doi: 10.1016/j.niox.2005.04.006

Singh, R., Hogg, N., Goss, S., Antholine, W., and Kalyanaraman, B. (1999). Mechanism of superoxide dismutase/H 2 O 2 -mediated nitric oxide release from S -nitrosoglutathione–role of glutamate. Arch. Biochem. Biophys. 372, 8–15. doi: 10.1006/abbi.1999.1447

Smith, N. G., and Dukes, J. S. (2013). Plant respiration and photosynthesis in global-scale models: incorporating acclimation to temperature and CO 2 . Glob. Change Biol. 19, 45–63. doi: 10.1111/j.1365-2486.2012.02797.x

Soden, B. J., Jackson, D. L., Ramaswamy, V., Schwarzkopf, M. D., and Huang, X. (2005). The radiative signature of upper tropospheric moistening. Science 310, 841–844. doi: 10.1126/science.1115602

Song, L., Ding, W., Zhao, M., Sun, B., and Zhang, L. (2006). Nitric oxide protects against oxidative stress under heat stress in the calluses from two ecotypes of reed. Plant Sci. 171, 449–458. doi: 10.1016/j.plantsci.2006.05.002

Song, Y., Miao, Y., and Song, C. P. (2014). Behind the scenes: the roles of reactive oxygen species in guard cells. New Phytol. 201, 1121–1140. doi: 10.1111/nph.12565

Stockwell, C., Hendry, A., and Kinnison, M. (2003). Contemporary evolution meets conservation. Trends Ecol. Evol. 18, 94–101. doi: 10.1016/S0169-5347(02)00044-7

Tossi, V., Lamattina, L., and Cassia, R. (2009). An increase in the concentration of abscisic acid is critical for nitric oxide-mediated plant adaptive responses to UV-B irradiation. New Phytol. 181, 871–879. doi: 10.1111/j.1469-8137.2008.02722.x

Uchida, A., Jagendorf, A. T., Hibino, T., Takabe, T., and Takabe, T. (2002). Effects of hydrogen peroxide and nitric oxide on both salt and heat stress tolerance in rice. Plant Sci. 163, 515–523. doi: 10.1016/S0168-9452(02)00159-0

Vainonen, J. P., and Kangasjärvi, J. (2015). Plant signalling in acute ozone exposure. Plant Cell Environ. 38, 240–252. doi: 10.1111/pce.12273

Voesenek, L. A., and Bailey-Serres, J. (2015). Flood adaptive traits and processes: an overview. New Phytol. 206, 57–73. doi: 10.1111/nph.13209

Wang, H., Xiao, W., Niu, Y., Chai, R., Jin, C., and Zhang, Y. (2015). Elevated carbon dioxide induces stomatal closure of Arabidopsis thaliana (L.) heynh. through an increased production of nitric oxide. J. Plant Growth Regul. 34, 372–380. doi: 10.1007/s00344-014-9473-6

Wuebbles, D. J., and Hayhoe, K. (2002). Atmospheric methane and global change. Earth Sci. Rev. 57, 177–210. doi: 10.1016/S0012-8252(01)00062-9

Xie, Y., Mao, Y., Zhang, W., Lai, D., Wang, Q., and Shen, W. (2014). Reactive oxygen species-dependent nitric oxide production contributes to hydrogen-promoted stomatal closure in Arabidopsis. Plant Physiol. 165, 759–773. doi: 10.1104/pp.114.237925

Xu, Z., Jiang, Y., Jia, B., and Zhou, G. (2016). Elevated-CO 2 response of stomata and its dependence on environmental factors. Front. Plant Sci. 7:657. doi: 10.3389/fpls.2016.00657

Xu, Z., Jiang, Y., and Zhou, G. (2015). Response and adaptation of photosynthesis, respiration, and antioxidant systems to elevated CO 2 with environmental stress in plants. Front. Plant Sci. 6:701. doi: 10.3389/fpls.2015.00701

Xue, L., Li, S., Sheng, H., Feng, H., Xu, S., and An, L. (2007). Nitric oxide alleviates oxidative damage induced by enhanced ultraviolet-B radiation in cyanobacterium. Curr. Microbiol. 55, 294–301. doi: 10.1007/s00284-006-0621-5

Yan, W., Zhong, Y., and Shangguan, Z. (2017). Contrasting responses of leaf stomatal characteristics to climate change: a considerable challenge to predict carbon and water cycles. Glob. Chang. Biol. 23, 3781–3793. doi: 10.1111/gcb.13654

Yang, H., Mu, J., Chen, L., Feng, J., Hu, J., Li, L., et al. (2015). S -Nitrosylation positively regulates ascorbate peroxidase activity during plant stress responses. Plant Physiol. 167, 1604–1615. doi: 10.1104/pp.114.255216

Yun, B.-W., Feechan, A., Yin, M., Saidi, N. B., Le Bihan, T., Yu, M., et al. (2011). S-nitrosylation of NADPH oxidase regulates cell death in plant immunity. Nature 478, 264–268. doi: 10.1038/nature10427

Zhang, Y., Tan, J., Guo, Z., Lu, S., He, S., Shu, W., et al. (2009). Increased abscisic acid levels in transgenic tobacco over-expressing 9 cis-epoxycarotenoid dioxygenase influence H 2 O 2 and NO production and antioxidant defenses. Plant Cell Environ. 32, 509–519. doi: 10.1111/j.1365-3040.2009.01945.x

Zinta, G., AbdElgawad, H., Domagalska, M. A., Vergauwen, L., Knapen, D., Nijs, I., et al. (2014). Physiological, biochemical, and genome-wide transcriptional analysis reveals that elevated CO 2 mitigates the impact of combined heat wave and drought stress in Arabidopsis thaliana at multiple organizational levels. Glob. Chang. Biol. 20, 3670–3685. doi: 10.1111/gcb.12626

Keywords : climate change, greenhouse effect, oxidative stress, nitric oxide, plants

Citation: Cassia R, Nocioni M, Correa-Aragunde N and Lamattina L (2018) Climate Change and the Impact of Greenhouse Gasses: CO 2 and NO, Friends and Foes of Plant Oxidative Stress. Front. Plant Sci. 9:273. doi: 10.3389/fpls.2018.00273

Received: 18 November 2017; Accepted: 16 February 2018; Published: 01 March 2018.

Reviewed by:

Copyright © 2018 Cassia, Nocioni, Correa-Aragunde and Lamattina. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY) . The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Raúl Cassia, [email protected]

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.

November 3, 2023

Earth Reacts to Greenhouse Gases More Strongly Than We Thought

Climate scientists, including pioneer James Hansen, are pinning down a fundamental factor that drives how hot Earth will get

By Chelsea Harvey & E&E News

Satellite image of Earth on black.

A 'Blue Marble' image of the Earth taken from the VIIRS instrument aboard NASA's Earth-observing satellite – Suomi NPP.

NASA/NOAA/GSFC/Suomi NPP/VIIRS/Norman Kuring

CLIMATEWIRE |  Climate scientist James Hansen is frustrated. And he’s worried.

For nearly 40 years, Hansen has been warning the world of the dangers of global warming. His testimony at a groundbreaking 1988 Senate hearing on the greenhouse effect helped inject the coming climate crisis into the public consciousness. And it helped make him one of the most influential climate scientists in the world.

Hansen has spent several decades as director of NASA’s Goddard Institute for Space Studies, and now at 82, he directs Columbia University’s  Climate Science, Awareness and Solutions program .

On supporting science journalism

If you're enjoying this article, consider supporting our award-winning journalism by subscribing . By purchasing a subscription you are helping to ensure the future of impactful stories about the discoveries and ideas shaping our world today.

In the years since his seminal testimony, many of Hansen’s basic scientific predictions about the Earth’s climate future have come true. Greenhouse gas emissions have grown, and global temperatures have continued to rise. The world’s glaciers and ice sheets are melting and sea level rise is accelerating.

But Hansen has been disappointed with the scientific community’s response to some of his more recent projections about the future of the warming Earth, which some researchers have characterized as unrealistically dire.

In particular, he was discouraged by the response to a paper he published in 2016, suggesting catastrophic ice melt in Greenland and Antarctica, with widespread global effects, may be possible with relatively modest future warming.

Many researchers said such outcomes were unlikely. But Hansen described the paper as some of his most important work and a warning about the need for more urgent action.

Now he’s bracing himself for a similar reaction to his  latest paper , published Thursday morning.

“I expect the response to be characterized by scientific reticence,” he said in an email to E&E News.

The new paper, published in the research journal  Oxford Open Climate Change , addresses a central question in modern climate science: How much will the Earth warm in response to future carbon emissions? It’s a metric known as “climate sensitivity,” or how sensitive the planet is to greenhouse gases in the atmosphere.

Hansen’s findings suggest the planet may warm faster than previous estimates have indicated. And while some experts say it’s possible, others suggest that he’s taken the results too far.

In studies, scientists often tackle the climate sensitivity question by investigating how much the Earth would warm if atmospheric carbon dioxide concentrations doubled their preindustrial levels. Prior to the industrial era, global CO2 levels hovered around 280 parts per million, meaning a doubling would land around 560 ppm.

Today’s CO2 levels have already climbed above 400 ppm, giving the question a growing relevance.

Climate sensitivity is a difficult metric to estimate. It hinges on a wide variety of feedback loops in the Earth’s climate system, which can speed up or slow down the planet’s warming.

As the Earth’s reflective glaciers and ice sheets melt, for instance, the planet can absorb more sunlight and warm at a faster rate. Forests and other natural ecosystems may absorb different amounts of carbon as the planet warms. Different types of clouds can both speed up or slow down global warming, and it’s still unclear how they will change as the Earth heats up.

The uncertainties around these factors have made it challenging for scientists to pin down an exact estimate for climate sensitivity. But they’ve chipped away at it in recent years.

For decades, studies generally suggested that the Earth should experience anywhere from 1.5 to 4.5 degrees Celsius of warming with a doubling of CO2. But a  2020 paper narrowed the range  to between 2.6 and 3.9 C, using multiple lines of evidence including climate models, the Earth’s response to recent historical emissions and the Earth’s ancient climate history.

The latest assessment report from the U.N.’s Intergovernmental Panel on Climate Change adopted a similar estimate, suggesting a likely range of 2.5 to 4 C with a central estimate around 3 C.

Hansen’s new paper, published with an international group of co-authors, significantly ups the numbers. It suggests a central estimate of around 4.8 C, nearly 2 degrees higher than the IPCC’s figure.

The paper relies largely on evidence from Earth’s ancient climate history. One reason? It’s unclear whether current climate models accurately represent all the relevant feedback effects that may affect climate sensitivity, Hansen and his co-authors argue. The planet’s past provides a clearer view of how the Earth has responded to previous shifts in atmospheric carbon dioxide concentrations.

The paper also suggests that global warming is likely to proceed faster in the near term than previous studies have suggested.

Under the international Paris climate agreement, world leaders are striving to keep global warming well below 2 C and below 1.5 C if at all possible. The new paper warns that warming could exceed 1.5 C by the end of the 2020s and 2 C by 2050.

A gradual global decline in air pollution, driven by tightening environmental regulations, is part of the reasoning. Some types of air pollution are known to have a cooling effect on the climate, which may mask some of the impact of greenhouse gas emissions. As these aerosols decline in the atmosphere, some research suggests, this masking effect may fall away and global temperatures may rise at faster rates.

Hansen and his co-authors argue that better accounting for the declines in global aerosols should accelerate estimates of near-term global warming. Studies suggest that warming between 1970 and 2010 likely proceeded at around 0.18 C per decade. Post-2010, the new paper argues, that figure should rise to 0.27 C.

The findings should motivate greater urgency to not only cut greenhouse gas emissions but to eventually lower global temperatures closer to their preindustrial levels, Hansen suggests. That means using natural resources and technological means to remove carbon dioxide from the atmosphere.

Hansen also suggests that a controversial form of geoengineering, known as solar radiation management, is likely warranted. SRM, in theory, would use reflective aerosols to beam sunlight away from the Earth and lower the planet’s temperatures. The practice has not been tested at any large scale, and scientists have raised a variety of concerns about its ethics and potential unintended side effects.

Yet Hansen believes scientists and activists “should raise concerns about the safety and ethics of NOT doing SRM,” he said by email.

Climate change, caused by human greenhouse gas emissions, is in itself a form of planetary geoengineering, he added.

“My suggestion is to reduce human geoengineering of the planet,” he said.

Yet some scientists say the new paper’s findings — again — are overblown.

The paper “adds very little to the literature,” said Piers Forster, director of the Priestly International Centre for Climate at Leeds University in the U.K. and a lead chapter author of the IPCC’s latest assessment report, in an email to E&E News.

It presents high-end estimates of climate sensitivity based on ancient climate records from the Earth’s past — but those findings aren’t necessarily new, he said. Forster also suggested that some of the methods the new paper used to arrive at those high estimates were “quite subjective and not justified by observations, model studies or literature.”

Forster also took issue with the new paper’s treatment of previous climate sensitivity estimates, including the widely cited 2020 study, which the authors suggested were far too low. The 2020 study presented a careful analysis, using multiple lines of high-quality evidence, Forster said. And yet the authors of the new paper “dismiss it, on spurious grounds.”

Michael Oppenheimer, a climate scientist and director of the Center for Policy Research on Energy and Environment at Princeton University, said the uncertainties around the effects of declining aerosols were important to pay attention to. And he suggested that the new paper’s climate sensitivity estimates were possible.

But added that he regards them as “a worst-worst-case” scenario.

“I think it’s perfectly legitimate to have a worst-worst-case out there,” he added. “They help people think about what the boundaries of the possible are, and they are necessary for risk management against the climate problem.”

But there are still so many uncertainties about the kinds of feedback factors affecting the Earth’s climate sensitivity, he said, that “you can’t really nail it down with the kind of precision that [Hansen’s] provided.”

But Hansen says the new paper’s lines of evidence are based on the most up-to-date research on the Earth’s ancient history.

“[T]here is no basis whatever for the claim that our results are ‘unlikely,’” he said by email. “It is the IPCC sensitivity that is unlikely, less than 1 percent chance of being right, as we show quantitatively in our (peer-reviewed) paper.”

Hansen and 'scientific reticence'

Hansen has been into the deep end of climate debates for much of his career.

In 1988, at the time of his Senate testimony, scientists were still discussing whether the fingerprint of human-caused global warming could yet be detected above the “noise” of the Earth’s natural climate variations.

“When I first got into this, and when Jim and I were testifying, we were arguing about whether there's a global signal,” said Oppenheimer, the Princeton scientist, who testified alongside Hansen in 1988. “All the information we had was about global mean temperature, global mean sea level. We couldn’t talk in the language of things that people cared about.”

But even with the limitations of climate science at the time, the scientists warned the world of the dangers to come.

Hansen has co-authored dozens of papers on climate change in the years since, many of which have been highly regarded by the scientific community.

“Over time, he’s got a pretty damn good track record of turning out to be right about things that other people thought differently about,” Oppenheimer said.

Forster, the Leeds University scientist, agreed that “some of Hansen’s papers are brilliant and his work and deeds helped establish this IPCC in the first place.”

But he added that he still thought the new paper misses the mark.

The reception is similar to a major paper Hansen published in 2016, widely known as the  “Ice Melt” paper.

The Ice Melt paper, published in the journal  Atmospheric Chemistry and Physics , provided a grim, sweeping vision of the Earth’s climate future, focused on the consequences of the melting Greenland and Antarctic ice sheets. Drawing largely on ancient climate data — similar to the new paper — it warned of rapid melting and sea-level rise on the order of several meters within the next century.

It also suggested that the rapid influx of cold, fresh meltwater into the sea could affect ocean circulation patterns and even cause a giant Atlantic current to shut down. That’s a controversial prediction  deemed unlikely by the IPCC , one that would have severe impacts on global weather and climate patterns if it actually happened.

The paper received mixed reactions from other climate scientists upon publication. Some praised the paper, while many suggested the findings were unrealistic.

Another 2016 paper , published by a different group of scientists, later found that the likelihood of an Atlantic current shutdown was relatively small and suggested that Hansen’s paper relied on “unrealistic assumptions.”

In his new paper, Hansen referred to that study as an “indictment” of Ice Melt. He also noted that the IPCC’s latest assessment report did not include Ice Melt’s predictions, an omission he likened in the new paper to a form of censorship.

“Science usually acknowledges alternative views and grants ultimate authority to nature,” the new paper states. “In the opinion of our first author (Hansen), IPCC does not want its authority challenged and is comfortable with gradualism. Caution has merits, but the delayed response and amplifying feedbacks of climate make excessive reticence a danger.”

Responding to critiques of his new paper, Hansen again suggested that “scientific reticence” — or a kind of resistance to new findings — is at play. He pointed to a  1961 paper by sociologist Bernard Barber  suggesting that scientists themselves can be resistant to scientific discovery.

Claims that his new findings are unrealistic, Hansen said, are “a perfect example of the category of scientific reticence that Barber describes as ‘resistance to discovery.’ It takes a long time for new results to sink into the community.”

Resistance to scientific findings is nothing new to Hansen. His 1988 testimony initially shook the political establishment — yet decades later, global climate action is still proceeding too slowly to meet the Paris climate targets.

When he first testified to Congress in the 1980s, Oppenheimer said, he expected that world governments would have started meaningful emissions reduction programs by the year 2000 or so.

“We didn’t get ahead of the impacts,” he said. “And that’s probably because people weren't willing to support strong governmental action in most countries … until they were getting clobbered by unusual and highly damaging, and in some cases unprecedented, climate events.”

He regards the current state of global climate action now with a mix of skepticism and optimism.

“We’re in the process of muddling through — we’re in a period where climate change is gonna be painful for a while, it’s gonna hurt a lot of people in a lot of places, but we can get out the other side,” he said. “I think we can get there. But will we?”

Hansen echoed his sentiments in starker terms.

He wrote that he’s been surprised by “the increase of anti-science no-nothing thinking in our politics.”

“That's why I focus on young people,” he added. “They need to understand the situation and take control.”

Reprinted from E&E News with permission from POLITICO, LLC. Copyright 2022. E&E News provides essential news for energy and environment professionals.

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Published: 14 December 1989

Observational determination of the greenhouse effect

  • A. Raval 1 &
  • V. Ramanathan 1  

Nature volume  342 ,  pages 758–761 ( 1989 ) Cite this article

3375 Accesses

381 Citations

44 Altmetric

Metrics details

Satellite measurements are used to quantify the atmospheric greenhouse effect, defined here as the infrared radiation energy trapped by atmospheric gases and clouds. The greenhouse effect is found to increase significantly with sea surface temperature. The rate of increase gives compelling evidence for the positive feedback between surface temperature, water vapour and the green-house effect; the magnitude of the feedback is consistent with that predicted by climate models. This study demonstrates an effective method for directly monitoring, from space, future changes in the greenhouse effect.

This is a preview of subscription content, access via your institution

Access options

Subscribe to this journal

Receive 51 print issues and online access

185,98 € per year

only 3,65 € per issue

Buy this article

  • Purchase on SpringerLink
  • Instant access to full article PDF

Prices may be subject to local taxes which are calculated during checkout

Similar content being viewed by others

research paper on greenhouse effect

Global warming at near-constant tropospheric relative humidity is supported by observations

research paper on greenhouse effect

Anthropogenic forcing and response yield observed positive trend in Earth’s energy imbalance

research paper on greenhouse effect

The residence time of water vapour in the atmosphere

Hansen, J. et al. Climate Processes and Climate Sensitivity (eds Hansen, J. & Takahashi, T.) 130–163 (Am. geophys. Un., Washington, DC, 1984).

Google Scholar  

Manabe, S. & Wetherald, R. T. J. atmos. Sci. 24 , 241–259 (1967).

Article   ADS   CAS   Google Scholar  

Cess, R. D. J. atmos. Sci. 33 , 1831–1843 (1976).

Article   ADS   Google Scholar  

Barkstrom, B. R. Bull. Am. met. Soc. 65 , 1170–1185 (1984).

Article   Google Scholar  

Ramanathan, V. et al. Science 243 , 57–62 (1989).

Beuttner, K. J. K. & Kern, C. D. J. geophys. Res. 70 , 1329–1337 (1965).

Reynolds, R. W. J. clim. 1 , 75–86 (1988).

Dickinson, R. E. & Cicerone, R. J. Nature 319 , 109–114 (1986).

Prabhakara, C., Short, D. A. & Vollmer, B. E. J. Clim. appl. Met. 24 , 1311–1324 (1985).

Mitchell, J. F. B., Wilson, C. A. & Cunnington, W. M. Q. Jl R. met. Soc. , 113 , 293–322 (1987).

Stone, P. H. & Carlson, J. H. J. atmos. Sci. 36 , 415–423 (1979).

Coakley, J. A. J. atmos. Sci. 34 , 465–470 (1977).

Prabhakara, C. Chang, H. D. & Chang, A. T. C. J. appl. Met. 21 , 59–68 (1982).

Stephens, G. J. Clim. (in the press).

Goody, R. M. Atmospheric Radiation Vol. 1 Ch. 9 (Oxford University Press, 1964).

Luther, F. M. World Climate Programme Vol. 93 (WM0, Geneva 1984).

Williamson, et al. NCAR Technical Note NCAR/TN-285 +STR (NCAR, Boulder, Colorado, 1987).

Kiehl, J. T. & Ramanathan, V. J. geophys. Res. (in the press).

Wetherald, R. T. & Manabe, S. J. atmos. Sci. 32 , 2044–2059 (1975).

Gadgil, S., Joseph, P. V. & Joshi, N. V. Nature 312 , 142–143 (1984).

Download references

Author information

Authors and affiliations.

Department of Geophysical Sciences, University of Chicago, Chicago, Illinois, 60637, USA

A. Raval & V. Ramanathan

You can also search for this author in PubMed   Google Scholar

Rights and permissions

Reprints and permissions

About this article

Cite this article.

Raval, A., Ramanathan, V. Observational determination of the greenhouse effect. Nature 342 , 758–761 (1989). https://doi.org/10.1038/342758a0

Download citation

Received : 21 August 1989

Accepted : 31 October 1989

Issue Date : 14 December 1989

DOI : https://doi.org/10.1038/342758a0

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

A new hybrid observation gnss tomography method combining the real and virtual inverted signals.

  • Wenyuan Zhang
  • Shubi Zhang
  • Xiaoming Wang

Journal of Geodesy (2021)

An earth system model shows self-sustained thawing of permafrost even if all man-made GHG emissions stop in 2020

  • Jorgen Randers
  • Ulrich Goluke

Scientific Reports (2020)

Deep learning and regression modelling of cloudless downward longwave radiation

  • Nsikan I. Obot
  • Ibifubara Humphrey
  • Sunday O. Udo

Beni-Suef University Journal of Basic and Applied Sciences (2019)

Electrochemical reduction of CO2 on Ni (OH)2 doped water dispersible graphene under different electrolyte conditions

  • Siva Palanisamy
  • Surendhiran Srinivasan

SN Applied Sciences (2019)

Convection and Climate: What Have We Learned from Simple Models and Simplified Settings?

  • Dennis L. Hartmann
  • Peter N. Blossey
  • Brittany D. Dygert

Current Climate Change Reports (2019)

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

research paper on greenhouse effect

National Academies Press: OpenBook

Climate Change: Evidence and Causes: Update 2020 (2020)

Chapter: conclusion, c onclusion.

This document explains that there are well-understood physical mechanisms by which changes in the amounts of greenhouse gases cause climate changes. It discusses the evidence that the concentrations of these gases in the atmosphere have increased and are still increasing rapidly, that climate change is occurring, and that most of the recent change is almost certainly due to emissions of greenhouse gases caused by human activities. Further climate change is inevitable; if emissions of greenhouse gases continue unabated, future changes will substantially exceed those that have occurred so far. There remains a range of estimates of the magnitude and regional expression of future change, but increases in the extremes of climate that can adversely affect natural ecosystems and human activities and infrastructure are expected.

Citizens and governments can choose among several options (or a mixture of those options) in response to this information: they can change their pattern of energy production and usage in order to limit emissions of greenhouse gases and hence the magnitude of climate changes; they can wait for changes to occur and accept the losses, damage, and suffering that arise; they can adapt to actual and expected changes as much as possible; or they can seek as yet unproven “geoengineering” solutions to counteract some of the climate changes that would otherwise occur. Each of these options has risks, attractions and costs, and what is actually done may be a mixture of these different options. Different nations and communities will vary in their vulnerability and their capacity to adapt. There is an important debate to be had about choices among these options, to decide what is best for each group or nation, and most importantly for the global population as a whole. The options have to be discussed at a global scale because in many cases those communities that are most vulnerable control few of the emissions, either past or future. Our description of the science of climate change, with both its facts and its uncertainties, is offered as a basis to inform that policy debate.

A CKNOWLEDGEMENTS

The following individuals served as the primary writing team for the 2014 and 2020 editions of this document:

  • Eric Wolff FRS, (UK lead), University of Cambridge
  • Inez Fung (NAS, US lead), University of California, Berkeley
  • Brian Hoskins FRS, Grantham Institute for Climate Change
  • John F.B. Mitchell FRS, UK Met Office
  • Tim Palmer FRS, University of Oxford
  • Benjamin Santer (NAS), Lawrence Livermore National Laboratory
  • John Shepherd FRS, University of Southampton
  • Keith Shine FRS, University of Reading.
  • Susan Solomon (NAS), Massachusetts Institute of Technology
  • Kevin Trenberth, National Center for Atmospheric Research
  • John Walsh, University of Alaska, Fairbanks
  • Don Wuebbles, University of Illinois

Staff support for the 2020 revision was provided by Richard Walker, Amanda Purcell, Nancy Huddleston, and Michael Hudson. We offer special thanks to Rebecca Lindsey and NOAA Climate.gov for providing data and figure updates.

The following individuals served as reviewers of the 2014 document in accordance with procedures approved by the Royal Society and the National Academy of Sciences:

  • Richard Alley (NAS), Department of Geosciences, Pennsylvania State University
  • Alec Broers FRS, Former President of the Royal Academy of Engineering
  • Harry Elderfield FRS, Department of Earth Sciences, University of Cambridge
  • Joanna Haigh FRS, Professor of Atmospheric Physics, Imperial College London
  • Isaac Held (NAS), NOAA Geophysical Fluid Dynamics Laboratory
  • John Kutzbach (NAS), Center for Climatic Research, University of Wisconsin
  • Jerry Meehl, Senior Scientist, National Center for Atmospheric Research
  • John Pendry FRS, Imperial College London
  • John Pyle FRS, Department of Chemistry, University of Cambridge
  • Gavin Schmidt, NASA Goddard Space Flight Center
  • Emily Shuckburgh, British Antarctic Survey
  • Gabrielle Walker, Journalist
  • Andrew Watson FRS, University of East Anglia

The Support for the 2014 Edition was provided by NAS Endowment Funds. We offer sincere thanks to the Ralph J. and Carol M. Cicerone Endowment for NAS Missions for supporting the production of this 2020 Edition.

F OR FURTHER READING

For more detailed discussion of the topics addressed in this document (including references to the underlying original research), see:

  • Intergovernmental Panel on Climate Change (IPCC), 2019: Special Report on the Ocean and Cryosphere in a Changing Climate [ https://www.ipcc.ch/srocc ]
  • National Academies of Sciences, Engineering, and Medicine (NASEM), 2019: Negative Emissions Technologies and Reliable Sequestration: A Research Agenda [ https://www.nap.edu/catalog/25259 ]
  • Royal Society, 2018: Greenhouse gas removal [ https://raeng.org.uk/greenhousegasremoval ]
  • U.S. Global Change Research Program (USGCRP), 2018: Fourth National Climate Assessment Volume II: Impacts, Risks, and Adaptation in the United States [ https://nca2018.globalchange.gov ]
  • IPCC, 2018: Global Warming of 1.5°C [ https://www.ipcc.ch/sr15 ]
  • USGCRP, 2017: Fourth National Climate Assessment Volume I: Climate Science Special Reports [ https://science2017.globalchange.gov ]
  • NASEM, 2016: Attribution of Extreme Weather Events in the Context of Climate Change [ https://www.nap.edu/catalog/21852 ]
  • IPCC, 2013: Fifth Assessment Report (AR5) Working Group 1. Climate Change 2013: The Physical Science Basis [ https://www.ipcc.ch/report/ar5/wg1 ]
  • NRC, 2013: Abrupt Impacts of Climate Change: Anticipating Surprises [ https://www.nap.edu/catalog/18373 ]
  • NRC, 2011: Climate Stabilization Targets: Emissions, Concentrations, and Impacts Over Decades to Millennia [ https://www.nap.edu/catalog/12877 ]
  • Royal Society 2010: Climate Change: A Summary of the Science [ https://royalsociety.org/topics-policy/publications/2010/climate-change-summary-science ]
  • NRC, 2010: America’s Climate Choices: Advancing the Science of Climate Change [ https://www.nap.edu/catalog/12782 ]

Much of the original data underlying the scientific findings discussed here are available at:

  • https://data.ucar.edu/
  • https://climatedataguide.ucar.edu
  • https://iridl.ldeo.columbia.edu
  • https://ess-dive.lbl.gov/
  • https://www.ncdc.noaa.gov/
  • https://www.esrl.noaa.gov/gmd/ccgg/trends/
  • http://scrippsco2.ucsd.edu
  • http://hahana.soest.hawaii.edu/hot/
was established to advise the United States on scientific and technical issues when President Lincoln signed a Congressional charter in 1863. The National Research Council, the operating arm of the National Academy of Sciences and the National Academy of Engineering, has issued numerous reports on the causes of and potential responses to climate change. Climate change resources from the National Research Council are available at .
is a self-governing Fellowship of many of the world’s most distinguished scientists. Its members are drawn from all areas of science, engineering, and medicine. It is the national academy of science in the UK. The Society’s fundamental purpose, reflected in its founding Charters of the 1660s, is to recognise, promote, and support excellence in science, and to encourage the development and use of science for the benefit of humanity. More information on the Society’s climate change work is available at

Image

Climate change is one of the defining issues of our time. It is now more certain than ever, based on many lines of evidence, that humans are changing Earth's climate. The Royal Society and the US National Academy of Sciences, with their similar missions to promote the use of science to benefit society and to inform critical policy debates, produced the original Climate Change: Evidence and Causes in 2014. It was written and reviewed by a UK-US team of leading climate scientists. This new edition, prepared by the same author team, has been updated with the most recent climate data and scientific analyses, all of which reinforce our understanding of human-caused climate change.

Scientific information is a vital component for society to make informed decisions about how to reduce the magnitude of climate change and how to adapt to its impacts. This booklet serves as a key reference document for decision makers, policy makers, educators, and others seeking authoritative answers about the current state of climate-change science.

READ FREE ONLINE

Welcome to OpenBook!

You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

Do you want to take a quick tour of the OpenBook's features?

Show this book's table of contents , where you can jump to any chapter by name.

...or use these buttons to go back to the previous chapter or skip to the next one.

Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

Switch between the Original Pages , where you can read the report as it appeared in print, and Text Pages for the web version, where you can highlight and search the text.

To search the entire text of this book, type in your search term here and press Enter .

Share a link to this book page on your preferred social network or via email.

View our suggested citation for this chapter.

Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

Get Email Updates

Do you enjoy reading reports from the Academies online for free ? Sign up for email notifications and we'll let you know about new publications in your areas of interest when they're released.

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

plants-logo

Article Menu

research paper on greenhouse effect

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Impact of plant community diversity on greenhouse gas emissions in riparian zones.

research paper on greenhouse effect

1. Introduction

2. materials and methods, 2.1. study site, 2.2. greenhouse gas (ghg) emission and global warming potential measurement, 2.3. soil samples collection and soil parameters measurement, 2.4. statistical analysis, 3.1. plant community diversity and soil parameters, 3.2. soil ghg emissions and global warming potential, 3.3. response of soil ghg emissions to soil parameter, 3.4. influence path of plant community diversity on soil ghg emissions, 4. discussion, 4.1. direct effect of plant community landscape diversity on soil ghg emissions, 4.2. plant community landscape diversity altered ghg emissions through changing soil parameters, 4.3. implications for managing plant community diversity to mitigate soil ghg emissions in riparian zones, 5. conclusions, author contributions, data availability statement, acknowledgments, conflicts of interest.

  • IPCC. Climate change 2021: The physical science basis. In Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change ; Zhai, V.P., Pirani, A., Connors, S.L., Péan, C., Berger, S., Caud, N., Chen, Y., Goldfarb, L., Gomis, M.I., Huang, M., et al., Eds.; Cambridge University Press: Cambridge, UK, 2021. [ Google Scholar ]
  • Cao, X.; Tang, S.; Liu, H.; Wen, L.; Kou, X.; Yu, X.; Wang, Q.; Wang, J.; Liu, D.; Zhuo, Y.; et al. Effects of plant communities on the emission of soil greenhouse gases in riparian wetlands during spring thaw. Ecohydrology 2022 , 10 , 1002. [ Google Scholar ] [ CrossRef ]
  • Hooper, D.; Adair, E.; Cardinale, B.; Byrnes, J.; Hungate, B.; Matulich, K.; Gonzalez, A.; Duffy, J.; Gamfeldt, L.; O’Connor, M. A global synthesis reveals biodiversity loss as a major driver of ecosystem change. Nature 2012 , 10 , 1038. [ Google Scholar ] [ CrossRef ]
  • Cui, M.; Wang, J.; Zhang, X.; Wang, C.; Li, G.; Wan, J.S.H.; Du, D. Warming significantly inhibited the competitive advantage of native plants in interspecific competition under phosphorus deposition. Plant Soil 2023 , 486 , 503–518. [ Google Scholar ] [ CrossRef ]
  • Nazir, M.; Li, G.; Nazir, M.; Zulfiqar, F.; Siddique, K.; Iqbal, B.; Du, D. Harnessing soil carbon sequestration to address climate change challenges in agriculture. Soil Till. Res. 2024 , 237 , 105–959. [ Google Scholar ] [ CrossRef ]
  • Niklaus, P.A.; Le Roux, X.; Poly, F.; Buchmann, N.; Scherer-Lorenzen, M.; Weigelt, A.; Barnard, R.L. Plant species diversity affects soil–atmosphere fluxes of methane and nitrous oxide. Oecologia 2016 , 10 , 1007. [ Google Scholar ] [ CrossRef ]
  • Hsieh, S.H.; Yuan, C.S.; Ie, I.R.; Yang, L.; Lin, H.J.; Hsueh, M.L. In-situ measurement of greenhouse gas emissions from a coastal estuarine wetland using a novel continuous monitoring technology: Comparison of indigenous and exotic plant species. J. Environ. Manag. 2021 , 281 , 111–905. [ Google Scholar ] [ CrossRef ]
  • Marín-Muñiz, J.L.; Hernandez, M.E.; Moreno-Casasola, P. Greenhouse gas emissions from coastal fresh water wetlands in Veracruz Mexico: Effect of plant community and seasonal dynamics. Atmos. Environ. 2015 , 107 , 107–117. [ Google Scholar ] [ CrossRef ]
  • Vroom, R.J.E.; van den Berg, M.; Pangala, S.R.; van der Scheer, O.E.; Sorrell, B.K. Physiological processes affecting methane transport by wetland vegetation—A review. Aquat. Bot. 2022 , 182 , 103–547. [ Google Scholar ] [ CrossRef ]
  • Yao, H.; Peng, H.; Hong, B.; Ding, H.; Hong, Y.; Zhu, Y.; Wang, J.; Cai, C. Seasonal and diurnal variation in ecosystem respiration and environmental controls from an alpine wetland in arid northwest China. J. Plant Ecol. 2022 , 15 , 933–946. [ Google Scholar ] [ CrossRef ]
  • Tao, F.; Huang, Y.; Hungate, B.; Manzoni, S.; Frey, S.; Schmidt, M.; Reichstein, M.; Carvalhais, N.; Ciais, P.; Jiang, L.; et al. Microbial carbon use efficiency promotes global soil carbon storage. Nature 2023 , 618 , 981–985. [ Google Scholar ] [ CrossRef ]
  • Buragienė, S.; Šarauskis, E.; Romaneckas, K.; Adamavičienė, A.; Kriaučiūnienė, Z.; Avižienytė, D.; Marozas, V.; Naujokienė, V. Relationship between CO 2 emissions and soil properties of differently tilled soils. Sci. Total Environ. 2019 , 662 , 786–795. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Khan, J.; Tariq, M.; Alabbosh, K.; Rehman, A.; Jalal, A.; Khan, A.; Farooq, M.; Li, G.; Iqbal, B.; Ahmad, N.; et al. Soil microplastics: Impacts on greenhouse gasses emissions, carbon cycling, microbial diversity, and soil characteristics. Appl. Soil Ecol. 2024 , 197 , 105–343. [ Google Scholar ] [ CrossRef ]
  • Han, W.; Shi, M.; Chang, J.; Ren, Y.; Xu, R.; Zhang, C.; Ge, Y. Plant species diversity reduces N 2 O but not CH 4 emissions from constructed wetlands under high nitrogen levels. Environ. Sci. Pollut. Res. 2017 , 10 , 1007. [ Google Scholar ] [ CrossRef ]
  • Wang, C.; Wu, B.; Jiang, K.; Wei, M.; Wang, S. Effects of different concentrations and types of Cu and Pb on soil N-fixing bacterial communities in the wheat rhizosphere. Appl. Soil Ecol. 2019 , 144 , 51–59. [ Google Scholar ] [ CrossRef ]
  • Khan, K.; Tang, Y.; Cheng, P.; Song, Y.; Li, X.; Lou, J.; Iqbal, B.; Zhao, X.; Hameed, R.; Li, G.; et al. Effects of degradable and non-degradable microplastics and oxytetracycline co-exposure on soil N 2 O and CO 2 emissions. Appl. Soil Ecol. 2024 , 197 , 105–331. [ Google Scholar ] [ CrossRef ]
  • Guenet, B.; Gabrielle, B.; Chenu, C.; Arrouays, D.; Balesdent, J.; Bernoux, M.; Bruni, E.; Caliman, J.; Cardinael, R.; Chen, S.; et al. Can N 2 O emissions offset the benefits from soil organic carbon storage? Glob. Change Biol. 2021 , 27 , 237–256. [ Google Scholar ] [ CrossRef ]
  • Philippot, L.; Hallin, S.; Borjesson, G.; Baggs, E.M. Biochemical cycling in the rhizosphere having an impact on global change. Plant Soil 2009 , 10 , 1007. [ Google Scholar ] [ CrossRef ]
  • Chang, J.; Fan, X.; Sun, H.; Zhang, C.; Song, C.; Chang, S.X.; Gu, B.; Liu, Y.; Li, D.; Wang, Y.; et al. Plant species richness enhances nitrous oxide emissions in microcosms of constructed wetlands. Ecol. Eng. 2014 , 64 , 108–115. [ Google Scholar ] [ CrossRef ]
  • Maucieri, C.; Borin, M.; Milani, M.; Cirelli, G.L.; Barbera, A.C. Plant species effect on CO 2 and CH 4 emissions from pilot constructed wetlands in Mediterranean area. Ecol. Eng. 2019 , 134 , 112–117. [ Google Scholar ] [ CrossRef ]
  • Mueller, P.; Mozdzer, T.J.; Langley, J.A.; Aoki, L.R.; Noyce, G.L.; Megonigal, J.P. Plant species determine tidal wetland methane response to sea level rise. Nat. Commun. 2020 , 11 , 5154. [ Google Scholar ] [ CrossRef ]
  • Chen, M.; Chang, L.; Zhang, J.; Guo, F.; Vymazal, J.; He, Q.; Chen, Y. Global nitrogen input on wetland ecosystem: The driving mechanism of soil labile carbon and nitrogen on greenhouse gas emissions. Environ. Sci. Ecotechnol. 2020 , 4 , 100063. [ Google Scholar ] [ CrossRef ]
  • Das, N.; Mandal, S. Microbial populations regulate greenhouse gas emissions in Sundarban mangrove ecosystem, India. Acta Ecol. Sin. 2022 , 42 , 641–652. [ Google Scholar ] [ CrossRef ]
  • Ding, H.; Liu, T.; Hu, Q.; Liu, M.; Cai, M.; Jiang, Y.; Cao, C. Effect of microbial community structures and metabolite profile on greenhouse gas emissions in rice varieties. Environ. Pollut. 2022 , 306 , 119–365. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Han, W.J.; Chang, J.; Fan, X.; Du, Y.Y.; Chang, S.X.; Zhang, C.B.; Ge, Y. Plant species diversity impacts nitrogen removal and nitrous oxide emissions as much as carbon addition in constructed wetland microcosms. Ecol. Eng. 2016 , 93 , 144–151. [ Google Scholar ] [ CrossRef ]
  • Zhao, Z.Y.; Chang, J.; Han, W.J.; Wang, M.; Ma, D.P.; Du, Y.Y.; Qu, Z.L.; Chang, S.X.; Ge, Y. Effects of plant diversity and sand particle size on methane emission and nitrogen removal in microcosms of constructed wetlands. Ecol. Eng. 2016 , 95 , 390–398. [ Google Scholar ] [ CrossRef ]
  • Luo, B.; Du, Y.Y.; Han, W.J.; Geng, Y.; Wang, Q.; Duan, Y.Y.; Ren, Y.; Liu, D.; Chang, J.; Ge, Y. Reduce health damage cost of greenhouse gas and ammonia emissions by assembling plant diversity in floating constructed wetlands treating wastewater. J. Clean. Prod. 2020 , 244 , 118–927. [ Google Scholar ] [ CrossRef ]
  • Wang, L.; Luo, B.; Du, Y.; Jiang, H.; Chang, S.X.; Fan, X.; Chang, J.; Ge, Y. Carbon nanotubes and plant diversity reduce greenhouse gas emissions and improve nitrogen removal efficiency of constructed wetlands. J. Clean. Prod. 2022 , 380 , 135023. [ Google Scholar ] [ CrossRef ]
  • Hu, Z.; Li, J.; Shi, K.; Ren, G.; Dai, Z.; Sun, J.; Zheng, X.; Zhou, Y.; Zhang, J.; Li, G.; et al. Effects of Canada goldenrod invasion on soil extracellular enzyme activities and ecoenzymatic stoichiometry. Sustainability 2021 , 13 , 3768. [ Google Scholar ] [ CrossRef ]
  • Hu, Z.; Zhang, J.; Du, Y.; Shi, K.; Ren, G.; Iqbal, B.; Dai, Z.; Li, G.; Du, D. Substrate availability regulates the suppressive effects of Canada goldenrod invasion on soil respiration. J. Plant Ecol. 2022 , 15 , 509–523. [ Google Scholar ] [ CrossRef ]
  • Xu, S.; Li, K.; Li, G.; Hu, Z.; Zhang, J.; Iqbal, B.; Du, D. Canada Goldenrod invasion regulates the effects of soil moisture on soil respiration. Int. J. Environ. Res. Public Health 2022 , 19 , 15446. [ Google Scholar ] [ CrossRef ]
  • Jiang, H.; Du, Y.; Han, W.; Wang, L.; Xiang, C.; Ge, Y.; Chang, J. Assembling plant diversity mitigates greenhouse gas emissions and achieves high nitrogen removal when treating the low-C/N wastewater by constructed wetlands. Environ. Sci. Pollut. R 2022 , 30 , 228–241. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wilsey, B.; Stirling, G. Species richness and evenness respond in a different manner to propagule density in developing prairie microcosm communities. Plant Ecol. 2007 , 190 , 259–273. [ Google Scholar ] [ CrossRef ]
  • Cheng, X.; Peng, R.; Chen, J.; Luo, Y.; Zhang, Q.; An, S.; Chen, J.; Li, B. CH 4 and N 2 O emissions from Spartina alterniflora and Phragmites australis in experimental mesocosms. Chemosphere 2007 , 68 , 420–427. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • IPCC. Climate Change 2013: The Physical Science Basis. In Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change ; Cambridge University Press: Cambridge, UK; New York, NY, USA, 2013. [ Google Scholar ]
  • Li, G.; Tang, Y.; Son, Y.; Zhao, X.; Iqbal, B.; Khan, K.; Guo, R.; Yin, W.; Zhao, X.; Du, D. Divergent responses in microbial metabolic limitations and carbon use efficiency to variably sized polystyrene microplastics in soil. Land Degrad. Dev. 2024 , 35 , 2658–2671. [ Google Scholar ] [ CrossRef ]
  • Tang, Y.; Li, X.; Song, Y.; Alabbosh, K.; Li, K.; Lou, J.; Hameed, R.; Iqbal, B.; Li, G.; Du, D. Nitrogen deposition enhanced the effect of Solidago Canadensis invasion on soil microbial metabolic limitation and carbon use efficiency. Pol. J. Environ. Stud. 2023 , 10 , 15244. [ Google Scholar ] [ CrossRef ]
  • Tang, Y.; Li, G.; Iqbal, B.; Tariq, M.; Rehman, A.; Khan, I.; Du, D. Soil nutrient levels regulate the effect of soil microplastic contamination on microbial element metabolism and carbon use efficiency. Ecotox. Environ. Saf. 2023 , 267 , 115–640. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • R Core Team. R: A Language and Environment for Statistical Computing ; R Core Team: Vienna, Austria, 2013; Available online: https://www.R-project.org (accessed on 1 January 2013).
  • Schmid, M.W.; van Moorsel, S.J.; Hahl, T.; De Luca, E.; De Deyn, G.B.; Wagg, C.; Niklaus, P.; Schmid, B. Effects of plant community history, soil legacy and plant diversity on soil microbial communities. J. Ecol. 2021 , 109 , 3007–3023. [ Google Scholar ] [ CrossRef ]
  • Liu, X.; Guo, Y.; Hu, H.; Sun, C.; Zhao, X.; Wei, C. Dynamics and controls of CO 2 and CH 4 emissions in the wetland of a montane permafrost region, Northeast China. Atmos. Environ. 2015 , 122 , 454–462. [ Google Scholar ] [ CrossRef ]
  • Bansal, S.; Johnson, O.F.; Meier, J.; Zhu, X. Vegetation affects timing and location of wetland methane emissions. J. Geophys. Res. Biogeosci. 2020 , 125 , e2020JG005777. [ Google Scholar ] [ CrossRef ]
  • Ma, W.; Alhassan, A.R.M.; Wang, Y.; Li, G.; Wang, H.; Zhao, J. Greenhouse gas emissions as influenced by wetland vegetation degradation along a moisture gradient on the eastern Qinghai-Tibet Plateau of North-West China. Nutr. Cycl. Agroecosys. 2018 , 112 , 335–354. [ Google Scholar ] [ CrossRef ]
  • Li, G.; Kim, S.; Han, S.H.; Chang, H.; Son, Y. Effect of soil moisture on the response of soil respiration to open-field experimental warming and precipitation manipulation. Forests 2017 , 8 , 56. [ Google Scholar ] [ CrossRef ]
  • Duan, J.; Cao, P.; Shu, T.; Zhou, B.; Xue, L.; Yang, L. Effects of cellulosic carbon addition on nitrogen removal from simulated dry land drainage, and its environmental effects. Agronomy 2023 , 13 , 304–412. [ Google Scholar ] [ CrossRef ]
  • Zhang, G.; Zhou, G.; Zhou, X.; Zhou, L.; Shao, J.; Liu, R.; Gao, J.; He, Y.; Du, Z.; Tang, J. Effects of tree mycorrhizal type on soil respiration and carbon stock via fine root biomass and litter dynamic in tropical plantations. J. Plant Ecol. 2023 , 10 , 1093. [ Google Scholar ] [ CrossRef ]
  • Olefeldt, D.; Turetsky, M.; Crill, P.; David, M. Environmental and physical controls on northern terrestrial methane emissions across permafrost zones. Glob. Change Biol. 2013 , 10 , 1111. [ Google Scholar ] [ CrossRef ]
  • Fu, W.; Zhang, J.; Wang, D.; Li, P.; Yin, Q. Effects of soil moisture on Phragmites australis (cav.) allelochemicals in soil and on growth of Phalaris arundinacea L. in Chinese wetland. Allelopath. J. 2020 , 10 , 26651. [ Google Scholar ] [ CrossRef ]
  • Lin, Q.; Wang, S.; Li, Y.; Riaz, L.; Yu, F.; Yang, Q.; Han, J.; Ma, J. Effects and mechanisms of land-types conversion on greenhouse gas emissions in the Yellow River floodplain. wetland. Sci. Total Environ. 2022 , 813 , 152–406. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zhang, J.; Wang, J.J.; Xiao, R.; Deng, H.; DeLaune, R.D. Effect of salinity on greenhouse gas production and emission in marsh soils during the decomposition of wetland plants. J. Soils Sediments 2022 , 23 , 133–144. [ Google Scholar ] [ CrossRef ]
  • Li, G.; Kim, S.; Han, S.H.; Chang, H.; Du, D.; Son, Y. Precipitation affects soil microbial and extracellular enzymatic responses to warming. Soil Biol. Biochem. 2018 , 120 , 212–221. [ Google Scholar ] [ CrossRef ]
  • Zhang, L.; Wang, J.; Zhao, R.; Guo, Y.; Hao, L. Aboveground net primary productivity and soil respiration display different responses to precipitation changes in desert grassland. J. Plant Ecol. 2022 , 15 , 57–70. [ Google Scholar ] [ CrossRef ]
  • Voigt, C.; Marushchak, M.E.; Abbott, B.W.; Biasi, C.; Elberling, B.; Siciliano, S.D.; Sonnentag, O.; Stewart, K.J.; Yang, Y.; Martikainen, P.J. Nitrous oxide emissions from permafrost affected soils. Nat. Rev. Earth Environ. 2020 , 1 , 420–434. [ Google Scholar ] [ CrossRef ]
  • Diao, H.; Chen, X.; Wang, G.; Ning, Q.; Hu, S.; Sun, W.; Dong, K.; Wang, C. The response of soil respiration to different n compounds addition in a saline-alkaline grassland of northern China. J. Plant Ecol. 2022 , 15 , 897–910. [ Google Scholar ] [ CrossRef ]
  • Troxler, T.G.; Ikenaga, M.; Scinto, L.; Boyer, J.N.; Condit, R.; Perez, R.; Gann, G.D.; Childers, D.L. Patterns of soil bacteria and canopy community structure related to tropical peatland development. Wetlands 2012 , 32 , 769–782. [ Google Scholar ] [ CrossRef ]
  • Zhang, H.; Goncalves, P.; Copeland, E.; Qi, S.; Dai, Z.; Li, G.; Wang, C.; Du, D.; Thomas, T. Invasion by the weed Conyza canadensis alters soil nutrient supply and shifts microbiota structure. Soil. Biol. Biochem. 2020 , 143 , 107739. [ Google Scholar ] [ CrossRef ]
  • Sun, H.; Xin, Q.; Ma, Z.; Lan, S. Effects of plant diversity on carbon dioxide emissions and carbon removal in laboratory-scale constructed wetland. Environ. Sci. Pollut. Res. 2019 , 10 , 1007. [ Google Scholar ] [ CrossRef ]
  • Huo, L.; Zou, Y.; Lyu, X.; Zhang, Z.; Wang, X.; An, Y. Effect of wetland reclamation on soil organic carbon stability in peat mire soil around Xingkai Lake in Northeast China. Chin. Geogr. Sci. 2018 , 28 , 325–336. [ Google Scholar ] [ CrossRef ]

Click here to enlarge figure

ParametersFpSite 1Site 2Site 3Site 4
pH2.7ns8.30 ± 0.05 a8.18 ± 0.01 ab8.21 ± 0.05 ab8.15 ± 0.02 b
SM (w/w%)10.02**0.21 ± 0.01 b0.20 ± 0.01 b0.24 ± 0.01 a0.21 ± 0.01 b
IN (×10 μg N g soil)4.28*0.54 ± 0.07 b6.08 ± 0.44 ab2.53 ± 1.25 b11.63 ± 4.51 a
DOC (×10 mg C g soil)1.33ns0.25 ± 0.030.32 ± 0.040.27 ± 0.010.24 ± 0.025
DON (mg N g soil)3.65ns0.04 ± 0.010.24 ± 0.010.11 ± 0.060.46 ± 0.18
SAP (mg P g soil)2.12ns0.03 ± 0.010.10 ± 0.020.13 ± 0.030.13 ± 0.06
DO 3.57ns7.27 ± 2.701.33 ± 0.224.19 ± 1.640.71 ± 0.26
DO 4.81*10.13 ± 3.18 a3.29 ± 0.36 b2.22 ± 0.54 b2.60 ± 1.06 b
DO 5.39*1.55 ± 0.48 bc2.76 ± 0.84 ab0.79 ± 0.37 c3.52 ± 0.17 a
MBC (×10 mg C g soil)2.98ns0.83 ± 0.100.53 ± 0.080.66 ± 0.060.59 ± 0.053
MBN (×10 mg N g soil)1.03ns0.03 ± 0.020.37 ± 0.330.02 ± 0.020.06 ± 0.02
MBP (×10 mg P g soil)1.44ns0.03 ± 0.010.07 ± 0.010.04 ± 0.010.04 ± 0.02
MB (×10 )3.2ns3.75 ± 1.440.93 ± 0.458.84 ± 3.821.12 ± 0.30
MB (×10 )0.96ns3.22 ± 1.290.96 ± 0.411.64 ± 0.312.45 ± 1.43
MB 0.66ns1.13 ± 0.514.99 ± 4.070.68 ± 0.573.66 ± 2.91
EEA (×10 nmol h g soil)7.1*2.75 ± 1.33 b4.55 ± 2.06 b2.65 ± 1.46 b7.26 ± 1.59 a
EEA (×10 nmol h g soil)414.21***2.94 ± 0.02 c4.31 ± 0.07 b2.93 ± 0.12 c6.05 ± 0.04 a
EEA (×10 nmol h g soil)129.1***2.97 ± 0.02 b2.93 ± 0.01 b2.80 ± 0.02 c3.59 ± 0.05 a
VL (×10 )5.02*0.93 ± 0.05 a1.56 ± 0.06 ab0.95 ± 0.05 b2.04 ± 0.47 a
VA (°)101.59***5.78 ± 0.09 a3.88 ± 0.08 ab5.45 ± 0.19 b3.40 ± 0.07 a
The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Li, G.; Xu, J.; Tang, Y.; Wang, Y.; Lou, J.; Xu, S.; Iqbal, B.; Li, Y.; Du, D. Impact of Plant Community Diversity on Greenhouse Gas Emissions in Riparian Zones. Plants 2024 , 13 , 2412. https://doi.org/10.3390/plants13172412

Li G, Xu J, Tang Y, Wang Y, Lou J, Xu S, Iqbal B, Li Y, Du D. Impact of Plant Community Diversity on Greenhouse Gas Emissions in Riparian Zones. Plants . 2024; 13(17):2412. https://doi.org/10.3390/plants13172412

Li, Guanlin, Jiacong Xu, Yi Tang, Yanjiao Wang, Jiabao Lou, Sixuan Xu, Babar Iqbal, Yingnan Li, and Daolin Du. 2024. "Impact of Plant Community Diversity on Greenhouse Gas Emissions in Riparian Zones" Plants 13, no. 17: 2412. https://doi.org/10.3390/plants13172412

Article Metrics

Article access statistics, further information, mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

The Greenhouse Effect and our Planet

The greenhouse effect happens when certain gases, which are known as greenhouse gases, accumulate in Earth’s atmosphere. Greenhouse gases include carbon dioxide (CO 2 ), methane (CH 4 ), nitrous oxide (N 2 O), ozone (O 3 ), and fluorinated gases.

Biology, Ecology, Earth Science, Geography, Human Geography

Loading ...

Newsela

The greenhouse effect happens when certain gases , which are known as greenhouse gases , accumulate in Earth’s atmosphere . Greenhouse gases include carbon dioxide (CO 2 ), methane (CH 4 ), nitrous oxide (N 2 O), ozone (O 3 ), and fluorinated gases.

Greenhouse gases allow the sun’s light to shine onto Earth’s surface, and then the gases, such as ozone, trap the heat that reflects back from the surface inside Earth’s atmosphere. The gases act like the glass walls of a greenhouse—thus the name, greenhouse gas

According to scientists, the average temperature of Earth would drop from 14˚C (57˚F) to as low as –18˚C (–0.4˚F), without the greenhouse effect.

Some greenhouse gases come from natural sources, for example, evaporation  adds water vapor to the atmosphere. Animals and plants release carbon dioxide when they respire, or breathe. Methane is released naturally from decomposition. There is evidence that suggests methane is released in low-oxygen environments , such as  swamps or landfills . Volcanoes —both on land and under the ocean —release greenhouse gases, so periods of high volcanic activity tend to be warmer.

Since the  Industrial Revolution  of the late 1700s and early 1800s, people have been releasing larger quantities of greenhouse gases into the atmosphere. That amount has skyrocketed in the past century. Greenhouse gas emissions increased 70 percent between 1970 and 2004. Emissions of CO 2 , rose by about 80 percent during that time.

The amount of CO 2 in the atmosphere far exceeds the naturally occurring range seen during the last 650,000 years.

Most of the CO 2 that people put into the atmosphere comes from burning  fossil fuels . Cars, trucks, t rains , and planes all burn fossil fuels. Many electric power plants do as well. Another way humans release CO 2 into the atmosphere is by cutting down  forests , because trees contain large amounts of carbon.

People add methane to the atmosphere through  livestock  farming, landfills, and fossil fuel production such as  coal mining  and natural gas processing. Nitrous oxide comes from  agriculture  and fossil fuel burning. Fluorinated gases include chlorofluorocarbons (CFCs),  hydrochlorofluorocarbons (HCFCs), and hydrofluorocarbons (HFCs). They are produced during the manufacturing of refrigeration and cooling products and through aerosols.

All of these human activities add greenhouse gases to the atmosphere. As the level of these gases rises, so does the temperature of Earth. The rise in Earth’s average temperature contributed to by human activity is known as  global warming .

The Greenhouse Effect and Climate Change Even slight increases in average global temperatures can have huge effects.

Perhaps the biggest, most obvious effect is that  glaciers and  ice caps melt faster than usual. The  meltwater  drains into the oceans, causing  sea levels to rise.

Glaciers and ice caps cover about 10 percent of the world’s landmasses. They hold between 70 and 75 percent of the world’s  freshwater . If all of this ice melted, sea levels would rise by about 70 meters (230 feet).

The Intergovernmental Panel on Climate Change states that the global sea level rose about 1.8 millimeters (0.07 inches) per year from 1961 to 1993, and about 3.1 millimeters (0.12 inches) per year since 1993.

Rising sea levels cause  flooding in  coastal cities, which could displace millions of people in low-lying areas such as Bangladesh, the U.S. state of Florida, and the Netherlands.

Millions more people in countries like Bolivia, Peru, and India depend on glacial meltwater for drinking,  irrigation , and  hydroelectric power . Rapid loss of these glaciers would devastate those countries.

Greenhouse gas emissions affect more than just temperature. Another effect involves changes in  precipitation , such as rain and  snow .

Over the course of the 20th century, precipitation increased in eastern parts of North and South America, northern Europe, and northern and central Asia. However, it has decreased in parts of Africa, the Mediterranean, and southern Asia.

As climates change, so do the habitats for living things. Animals that are adapted to a certain climate may become threatened. Many human societies depend on predictable rain patterns in order to grow specific  crops for food, clothing, and trade. If the climate of an area changes, the people who live there may no longer be able to grow the crops they depend on for survival. Some scientists also worry that tropical diseases will expand their ranges into what are now more temperate regions if the temperatures of those areas increase.

Most climate scientists agree that we must reduce the amount of greenhouse gases released into the atmosphere. Ways to do this, include:

  • driving less, using public transportation , carpooling, walking, or riding a bike.
  • flying less—airplanes produce huge amounts of greenhouse gas emissions.
  • reducing, reusing, and recycling.
  • planting a tree—trees absorb carbon dioxide, keeping it out of the atmosphere.
  • using less  electricity .
  • eating less meat—cows are one of the biggest methane producers.
  • supporting alternative energy sources that don’t burn fossil fuels.

Artificial Gas

Chlorofluorocarbons (CFCs) are the only greenhouse gases not created by nature. They are created through refrigeration and aerosol cans.

CFCs, used mostly as refrigerants, are chemicals that were developed in the late 19th century and came into wide use in the mid-20th century.

Other greenhouse gases, such as carbon dioxide, are emitted by human activity, at an unnatural and unsustainable level, but the molecules do occur naturally in Earth's atmosphere.

Media Credits

The audio, illustrations, photos, and videos are credited beneath the media asset, except for promotional images, which generally link to another page that contains the media credit. The Rights Holder for media is the person or group credited.

Illustrators

Educator reviewer, last updated.

August 21, 2024

User Permissions

For information on user permissions, please read our Terms of Service. If you have questions about how to cite anything on our website in your project or classroom presentation, please contact your teacher. They will best know the preferred format. When you reach out to them, you will need the page title, URL, and the date you accessed the resource.

If a media asset is downloadable, a download button appears in the corner of the media viewer. If no button appears, you cannot download or save the media.

Text on this page is printable and can be used according to our Terms of Service .

Interactives

Any interactives on this page can only be played while you are visiting our website. You cannot download interactives.

Related Resources

IMAGES

  1. Why is it happening?

    research paper on greenhouse effect

  2. What is climate change and how will it affect the UK?

    research paper on greenhouse effect

  3. Greenhouse Effect

    research paper on greenhouse effect

  4. Carbon cycle and greenhouse effect

    research paper on greenhouse effect

  5. Global Warming

    research paper on greenhouse effect

  6. Greenhouse Effect Paragraph

    research paper on greenhouse effect

VIDEO

  1. Scallop Square Mini Album

  2. Quick Revision Paper 2: Greenhouse Effect & Global Warming 🌍🔥 PART 8

  3. CSEC Physics 2024 Paper 2 Question 3 Solution

  4. What is the Greenhouse Effect?

  5. What is the Greenhouse Effect and How Does It Impact Our Climate

  6. Understanding CO2: Part 2

COMMENTS

  1. Greenhouse Effect: Greenhouse Gases and Their Impact on ...

    The clear effect of the greenhouse gases is the. stable heating of Earth's atmosphere and surface, thus, global warming. The abi lity of certain gases, greenhouse gases, to be transparent to ...

  2. Climate Change and the Impact of Greenhouse Gasses: CO

    Life on Earth is possible thanks to greenhouse effect. Without it, temperature on Earth's surface would be around -19°C, instead of the current average of 14°C. Greenhouse effect is produced by greenhouse gasses (GHG) like water vapor, carbon dioxide (CO 2), methane (CH 4), nitrous oxides (N x O) and ozone (O 3). GHG have natural and ...

  3. CO2, the greenhouse effect and global warming: from the pioneering work

    Climate change is a major risk facing mankind. At the United Nations Climate Change Conference held in Paris at the end of last year, 195 countries agreed on a plan to reduce emissions of CO 2 and other greenhouse gases, aiming to limit global temperature increase to well below 2 °C (relative to pre-industrial climate, meaning a future warming of less than 1.4 °C because temperature had ...

  4. Comprehensive review: Effects of climate change and greenhouse gases

    The greenhouse effect is identified as a primary contributor to climate change. The Earth is uniquely capable of sustaining life within our solar system due to its distinct environmental characteristics, such as the presence of water, an atmosphere abundant in oxygen, and a surface temperature conducive to life (Xian et al., 2023).The residual solar energy that reaches the Earth contributes to ...

  5. Greenhouse gases emissions and global climate change: Examining the

    Human activities significantly increase the concentration of GHG in the atmosphere, intensifying the greenhouse effect and impacting the climate (Althor et al., 2016; Zheng et al., 2019).According to data from Table 1, provided by the Emission Database for Global Atmospheric Research (EDGAR) and developed by the European Commission's Joint Research Centre (Crippa et al., 2018), the main ...

  6. PDF The Greenhouse Effect: Science and Policy

    The greenhouse effect, despite all the controversy that surrounds the term, is actually one of the most well-established theories in atmospheric science. For example, with its dense CO2 atmosphere, Venus has temperatures near 700 K at its surface. Mars, with its very thin CO2 atmosphere, has temperatures of only 220 K.

  7. The Greenhouse Effect: Science and Policy

    Global warming from the increase in greenhouse gases has become a major scientific and political issue during the past decade. That infrared radiation is trapped by greenhouse gases and particles in a planetary atmosphere and that the atmospheric CO 2 level has increased by some 25 percent since 1850 because of fossil fuel combustion and land ...

  8. Climate Change and the Impact of Greenhouse Gasses: CO

    FIGURE 1. Simplified scheme showing greenhouse gasses (GHG) and their effects on plants. GHG (H 2 O vapor, clouds, CO 2, CH 4, N 2 O, and NO) have both natural and anthropogenic origin, contributing to greenhouse effect. Short-term effects of GHG increase is mainly CO 2 rise, that activates photosynthesis (PS) and inhibits stomatal opening (SO). Long-term effects of GHG increase are extreme ...

  9. Persistence of climate changes due to a range of greenhouse gases

    Carbon dioxide, methane, nitrous oxide, and other greenhouse gases increased over the course of the 20th century due to human activities. The human-caused increases in these gases are the primary forcing that accounts for much of the global warming of the past fifty years, with carbon dioxide being the most important single radiative forcing agent ().

  10. How a Stable Greenhouse Effect on Earth Is Maintained Under Global

    Journal of Geophysical Research: Atmospheres is an AGU journal publishing original research articles that advance and improve the understanding of atmospheric properties and processes. ... Search for more papers by this author. David Paynter, ... when the runaway greenhouse effect was expected to occur (Koll & Cronin, 2018), the feedback can ...

  11. The greenhouse effect: Damages, costs and abatement

    The buildup of so-called "greenhouse gases" in the atmosphere — CO2 in particular-appears to be having an adverse impact on the global climate. This paper briefly reviews current expectations with regard to physical and biological effects, their potential costs to society, and likely costs of abatement. For a "worst case" scenario it is impossible to assess, in economic terms, the ...

  12. Earth Reacts to Greenhouse Gases More Strongly Than We Thought

    His testimony at a groundbreaking 1988 Senate hearing on the greenhouse effect helped inject the coming climate crisis into the public consciousness. ... The new paper, published in the research ...

  13. Role of greenhouse gas in climate change**

    Here, I discuss the role of greenhouse gases such as carbon dioxide and water vapour in global warming, using a hierarchy of climate models with increasing complexity. 1. Introduction. During the last several decades, climate models have been used very extensively for predicting global climate change.

  14. The runaway greenhouse: implications for future climate change

    The ultimate climate emergency is a 'runaway greenhouse': a hot and water-vapour-rich atmosphere limits the emission of thermal radiation to space, causing runaway warming. Warming ceases only after the surface reaches approximately 1400 K and emits radiation in the near-infrared, where water is not a good greenhouse gas.

  15. PDF Observational determination of the greenhouse effect

    The greenhouse effect of the atmosphere, next to solar absorption, is the largest factor in maintaining the present climate. The TOA flux is given by the radiation transfer equation F= B(T,)-LA(x ...

  16. Greenhouse effect and climate change: scientific basis and overview

    The natural greenhouse effect arises due to some of the trace gases, called the greenhouse gases, which are nearly transparent to solar radiation but strongly absorb the infra-red radiation emitted by the Earth. This results in a warming of the Earth by about 30°C and makes it habitable.

  17. Climate Change: Evidence and Causes: Update 2020

    C ONCLUSION. This document explains that there are well-understood physical mechanisms by which changes in the amounts of greenhouse gases cause climate changes. It discusses the evidence that the concentrations of these gases in the atmosphere have increased and are still increasing rapidly, that climate change is occurring, and that most of ...

  18. Greenhouse effect

    The greenhouse effect occurs in the atmosphere, and is an essential part of How the Earth System Works. Click the image on the left to open the Understanding Global Change Infographic. Locate the greenhouse effect icon and identify other topics that cause changes to, or are affected by, the greenhouse effect.

  19. Impact of Plant Community Diversity on Greenhouse Gas Emissions in

    Plant community succession can impact greenhouse gas (GHG) emissions from the soil by altering the soil carbon and nitrogen cycles. However, the effects of community landscape diversity on soil GHG emissions have rarely been fully understood. Therefore, this study investigated how plant landscape diversity, structure type, and species composition, affect soil GHG emissions in a riparian zone ...

  20. Greenhouse Effect

    greenhouse effect. phenomenon where gases allow sunlight to enter Earth's atmosphere but make it difficult for heat to escape. greenhouse gas. gas in the atmosphere, such as carbon dioxide, methane, water vapor, and ozone, that absorbs solar heat reflected by the surface of the Earth, warming the atmosphere.

  21. History of the greenhouse effect

    The greenhouse effect is now commonly accepted by the scientific community, politicians and the general public. However, the misnomer 'greenhouse effect' has perpetuated, and there are a number of aspects of the effect which are poorly understood outside the atmospheric sciences. On such misconception is that greenhouse research is a recent ...

  22. PDF TEACHER BACKGROUND: THE GREENHOUSE EFFECT

    ( 21%), exert almost no greenhouse effect. Instead, the greenhouse effect comes from molecules that are more complex and much less common. Water vapor is the most important greenhouse gas, and carbon dioxide (CO 2) is the second-most important one. Methane, nitrous oxide, ozone and several other gases present in the atmosphere in small amounts ...

  23. Greenhouse Effect

    The greenhouse effect occurs through radiation energy from the sun in the form of ultraviolet, visible and near-infrared radiation that is captured by the planet's atmosphere to warm the planet's surface—it is essential to support life. If there is a substantial increase in any of the greenhouse gases, there are changes in weather patterns with implications on the climate.

  24. The Greenhouse Effect and our Planet

    The greenhouse effect happens when certain gases, which are known as greenhouse gases, accumulate in Earth's atmosphere. Greenhouse gases include carbon dioxide (CO 2), methane (CH 4), nitrous oxide (N 2 O), ozone (O 3), and fluorinated gases.. Greenhouse gases allow the sun's light to shine onto Earth's surface, and then the gases, such as ozone, trap the heat that reflects back from ...