ArticlePDF Available

Causes of reduced leaf-level photosynthesis during strong El Niño drought in a Central Amazon forest Amazon photosynthesis resilience to drought

Authors:

Figures

No caption available
… 
Content may be subject to copyright.
Accepted Article
This article has been accepted for publication and undergone full peer review but has not
been through the copyediting, typesetting, pagination and proofreading process, which may
lead to differences between this version and the Version of Record. Please cite this article as
doi: 10.1111/gcb.14293
This article is protected by copyright. All rights reserved.
DR. VICTOR ALEXANDRE HARDT FERREIRA SANTOS (Orcid ID : 0000-0002-3510-5035)
Article type : Primary Research Articles
Causes of reduced leaf-level photosynthesis during strong El Niño drought in a Central
Amazon forest
Amazon photosynthesis resilience to drought
Victor Alexandre Hardt Ferreira dos Santos1; Marciel José Ferreira2*; João Victor Figueiredo
Cardoso Rodrigues3; Maquelle Neves Garcia1; João Vitor Barbosa Ceron1; Bruce Walker
Nelson1; Scott Reid Saleska4
1 Environmental Dynamics Department, Brazil´s National Institute for Amazon Research,
Manaus, Amazonas, Brazil. (vichardt@hotmail.com; maquelleneves@gmail.com;
joaovitorceron@gmail.com; bnelsonbr@gmail.com)
2Department of Forest Sciences, Federal University of Amazonas, Manaus, Amazonas,
Brazil. (mjf.ufam@gmail.com)
3Center for Distance Education, Federal University of Amazonas, Manaus, Amazonas, Brazil.
(joao.ufam@gmail.com)
4Department of Ecology and Evolutionary Biology, University of Arizona, Tucson, Arizona,
USA. (saleska@email.arizona.edu)
*(corresponding author): Telephone: +55 (92) 33054042; Email: mjf.ufam@gmail.com
Keywords: Tropical forest, warming, drought stress, stomatal conductance, climate change,
chlorophyll fluorescence
Accepted Article
This article is protected by copyright. All rights reserved.
Primary Research Article
Abstract
Sustained drought and concomitant high temperature may reduce photosynthesis and cause
tree mortality. Possible causes of reduced photosynthesis include stomatal closure and
biochemical inhibition, but their relative roles are unknown in Amazon trees during strong
drought events. We assessed the effects of the recent (2015) strong El Niño drought on leaf-
level photosynthesis of Central Amazon trees via these two mechanisms. Through four
seasons of 2015, we measured leaf gas exchange, chlorophyll a fluorescence parameters,
chlorophyll concentration and nutrient content in leaves of 57 upper canopy and understory
trees of a lowland terra firme forest on well-drained infertile oxisol. Photosynthesis
decreased 28% in the upper canopy and 17% in understory trees during the extreme dry
season of 2015, relative to other 2015 seasons and was also lower than the climatically
normal dry season of the following non-El Niño year. Photosynthesis reduction under
extreme drought and high temperature in the 2015 dry season was related only to stomatal
closure in both upper canopy and understory trees, and not to chlorophyll a fluorescence
parameters, chlorophyll, or leaf nutrient concentration. The distinction is important because
stomatal closure is a transient regulatory response that can reverse when water becomes
available, whereas the other responses reflect more permanent changes or damage to the
photosynthetic apparatus. Photosynthesis decrease due to stomatal closure during the 2015
extreme dry season was followed two months later by an increase in photosynthesis as rains
returned, indicating a margin of resilience to one-off extreme climatic events in Amazonian
forests.
Accepted Article
This article is protected by copyright. All rights reserved.
Keywords: Tropical forest, warming, drought stress, stomatal conductance, climate change,
chlorophyll fluorescence
Introduction
Amazon forests significantly influence the global carbon cycle, storing ~120 PgC in
biomass and contributing 14% of the net annual carbon fixed by terrestrial photosynthesis
(Malhi et al., 2006; Malhi et al., 2008; Zhao & Running, 2010; Fauset et al., 2015). However,
this function is at risk due to drought events in the region, linked to global climate change
(Aragão et al., 2014; Brienen et al., 2015). Reduction in ecosystem scale photosynthesis and
widespread tree mortality during and after drought have been reported, suggesting that these
forests may be increasingly at the risk of shift from a carbon sink to a source (Bonal et al.,
2008; Phillips et al., 2009; Aragão et al., 2014; Gatti et al., 2014; Brienen et al., 2015;
Doughty et al., 2015; Bonal et al., 2016; Cavaleri et al., 2017).
Concerning seasonal drought, evidence suggests that gross ecosystem productivity of
Central Amazon forests is not limited by rainfall during a typical dry season, but instead
increases due to coordinated dry season leaf-flush and litterfall that shifts the composition of
the canopy towards younger leaves with greater photosynthetic capacity (Kitajima et al.,
1997; Doughty & Goulden, 2008; Restrepo-Coupe et al., 2013; Wagner et al., 2016; Wu et
al., 2016; Wagner et al., 2017). However, important remaining questions include how in situ
leaves of a given age respond to drought or heat stress, and how they respond under more
extreme or long-term droughts like those occurring during dry phases of El Niño and the
Atlantic Multidecadal Oscillation (Davidson et al., 2012; Bonal et al., 2016).
The effects of normal seasonality and of extreme drought events on leaf-level
photosynthetic processes of mature leaves are much debated. Some reports indicate no
seasonal change in photosynthesis at the leaf level during normal but prolonged (5 mo)
Accepted Article
This article is protected by copyright. All rights reserved.
seasonal drought (Domingues et al., 2014) and maintenance of growth and photosynthetic
capacity (despite increased mortality by xylem embolism) during 15y of a through-fall
exclusion experiment (Rowland et al., 2015a; Rowland et al., 2015b) in the eastern Amazon.
Tree growth rates over a 2 y period that included the extreme 2005 drought were no different
from other multi-year intervals that did not include droughts (Phillips et al., 2009). On the
other hand, leaf-level photosynthesis reduction has been reported during the normal
prolonged seasonal drought typical of southern Amazonia (Miranda et al., 2005), during the
extreme 2010 drought in a Bolivian forest (Doughty et al., 2015) and at Paracou, French
Guiana, during a pronounced dry season in 2008 (Stahl et al., 2013a). The variability between
individual trees responses (both inter and intra-specific) to drought also adds to the difficulty
of estimating ecosystem-scale drought effects when examining leaf-level photosynthetic
processes in tropical forests (Miranda et al., 2005; Stahl et al., 2013a). Critical to resolving
questions about the sensitivity of leaf-level photosynthesis to drought and to accurately
representing tropical forest response to climate change in Earth System models is a more
detailed understanding of the mechanisms involved in the leaf-level response to drought and
to high temperatures (Corlett, 2016).
The mechanism of leaf-level photosynthesis reduction during drought stress might be
stomatal closure (which limits CO2 assimilation), non-stomatal limitations to photosynthetic
function, or a combination of the two (Flexas & Medrano, 2002; Lloyd & Farquhar, 2008).
Under increased vapor pressure deficit from low humidity and/or high temperature, plants
close their stomata (Cunningham, 2004; Slot & Winter, 2017). This maintains leaf turgor and
water column integrity in xylem vessels at least up to a critical leaf water potential despite
low availability of soil water and high atmospheric evaporative demand (Bonal et al., 2000;
Flexas & Medrano, 2002; Chastain et al., 2014; Sperlich et al., 2015). Stomatal closure
protects against hydraulic failure but reduces photosynthesis, limits CO2 diffusion to the
Accepted Article
This article is protected by copyright. All rights reserved.
substomatal cavity and reduces leaf cooling under the high irradiance and high temperatures
concurrent with natural drought (Teskey et al., 2015).
During the stomatal closure, reduced use of energy for CO2 assimilation can
overexcite chlorophyll. Together with high leaf temperature from reduced evaporative
cooling, this may damage the integrity and functionality of photochemical processes in the
chloroplast membranes (Oukarroum et al., 2009; Desotgiu et al., 2012, Campos et al., 2014).
The distinction between reductions in photosynthesis due to stomatal closure alone, versus
reductions that follow damage to leaf biochemistry, are critically important in terms of their
implications for long-term drought response of the forest. Stomatal closure is transient and
may be reversed once the drought is past, but damage to leaf photosynthetic infrastructure
may be longer-lived, and require valuable carbon and water resources to repair or replace
damaged leaves (Flexas & Medrano, 2002; Lloyd & Farquhar, 2008).
In 2015, a widespread warming and extreme drought occurred over the Amazon forest
due to a strong El Niño phase in the El Niño-Southern Oscillation (ENSO) (Jiménez-Muñoz
et al., 2016). We used this natural event to assess the effects of drought on leaf-level
photosynthesis and on the mechanisms which control it and to thereby improve our
understanding of climate change impacts on tropical forests.
We hypothesize that drought reduces photosynthesis in mature leaves. We endeavor to
confirm this general hypothesis and test further key hypotheses about the mechanisms
regulating photosynthesis under drought and high maximum temperatures in a Central
Amazon forest. In particular, we test whether changes in photosynthesis are due only to
changes in stomatal conductance (brought on by drought and/or temperature-induced increase
in vapor pressure deficit), or also due to other changes in photosynthetic function (as
indicated by photosynthetic capacity, chlorophyll fluorescence, chlorophyll concentration and
nutrient content).
Accepted Article
This article is protected by copyright. All rights reserved.
We examined 57 trees cohabiting the forest upper canopy and the understory during
four consecutive seasons in 2015 to assess the effects of the strong El Niño on leaf-level
photosynthesis and its regulation. The El Niño drought was strong in only the 2015 dry
season. However, to be sure that we were not detecting typical seasonal differences between
the 2015 dry season and the other three seasons of 2015, we made a second set of dry season
measurements in 2016, which had normal seasonal rainfall and temperature.
Materials and Methods
Study site
The study site is a Central Amazon forest at the LBA (Large Scale Biosphere-
Atmosphere Experiment in Amazonia) K-34 tower (2.6° S, 60.2° W). The region is
characterized by low seasonal variation in air temperature (24 - 27°C monthly average), high
humidity (75% - 92% daily average) and moderately high precipitation (2200 mm annually)
(Araújo et al., 2002). The terrain is undulating with altitude ranging from 60 m in valleys to
120 m a.s.l. on plateaus. Plateau soil is predominantly well-drained clay oxisol while valleys
and lower slopes have poorly drained sandy spodosol (Luizão et al., 2004). The forest
structure is characterized by high tree density (626 trees ha-1), tree basal area (28-30 m2 ha-1)
and tree aboveground biomass (360 Mg ha-1) (Chambers et al., 2004; Vieira et al., 2004).
We measured leaf traits in 19 trees inhabiting the upper canopy layer and 38 trees in
the understory (Table S1), during four field campaigns in the wet, wet/dry transition, dry and
dry/wet transition seasons of 2015. Campaign mid-dates in 2015 were May 7, July 7,
September 13 and November 26, respectively. Extreme drought occurred prior to and during
the September campaign. To compare the 2015 extreme dry season to a normal dry season,
gas exchange measurements were also made in the dry season of 2016, with a mid-date on
Oct 17, 2016.
Accepted Article
This article is protected by copyright. All rights reserved.
The trees are all located on a plateau in the footprint of K-34 LBA tower. The upper
canopy layer was accessed from the tower and two nearby canopy walkways, each 30m long
and 25-30m above the ground. In the understory, the leaves were accessible from the ground.
Environmental conditions
Daily and monthly precipitation for 1998-2016 were obtained from products 3B42v7
and 3B43v7, respectively of the Tropical Rainfall Measuring Mission database (NASA,
2016). We used the daily precipitation to calculate a 30 d running sum prior to each day.
From this we obtained the Accumulated Daily Water Deficit as of the mid-date of each field
campaign. This is a modified version of the Maximum Climatological Water Deficit, or
MCWD, often used as a metric of drought intensity across Amazonia (Malhi et al., 2009;
Phillips et al., 2009; Lewis et al., 2011; Aragão et al., 2014; see Text S5). In addition to our
modified MCWD, we used the soil moisture from six depths (10, 20, 30, 40, 60 and 100 cm)
to calculate the Relative Extractable Water index (REW) (Granier et al., 1999; Wagner et al.,
2011; Stahl et al., 2013a).
The maximum monthly air temperature was from the Brazilian National Institute of
Meteorology (INMET Manaus station #81730, 3.13° S; 59.95° W). Anomalies for monthly
rainfall and maximum air temperature were calculated according to Saleska et al. (2007). To
address rainfall and temperature associations with ENSO intensity, these monthly anomalies
were correlated to the Multivariate ENSO Index (MEI) (Wolter & Timlin, 1998;
http://www.cdc.noaa.gov/people/klaus.wolter/MEI/table.html). Air temperature and relative
humidity for the four 2015 field campaigns were obtained from a sensor (HMP45C, Vaisala
Oyj, Finland) installed above the canopy at 51 m height on the K-34 tower. A data gap for
relative humidity occurred in July. Photosynthetic photon flux density sensors were installed
near the ground (1.3 m) and above the canopy (51 m) during 5-7 days in each campaign
Accepted Article
This article is protected by copyright. All rights reserved.
(MQS-B PAR sensor and ULM-500 logger, Heinz Walz, Germany). Soil moisture was
determined at the six depths with a profile probe (PR1, Delta-T Devices, UK). Soil moisture,
air temperature and relative humidity also were recorded during the 2016 dry season.
Leaf traits
For all leaf traits, we selected samples from mature, healthy, fully expanded leaves
with no signs of aging or senescence. To control for leaf age effects, in a parallel study of leaf
demography, we monitored the leaf age of our upper canopy trees with a precision of +/- 15
days. This allowed us to confidently select only mature-stage leaves in each season and each
year of the present study. We measured one leaf per tree in each season for gas exchange
parameters and three leaves per tree for the other traits. Gas-exchange and chlorophyll a
fluorescence parameters were obtained in vivo from attached leaves and branches. The same
leaves were then harvested for the other measurements.
Light-saturated photosynthetic rate (Asat), dark respiration (Rd), transpiration (E) and
stomatal conductance (gs) were measured between 08:00h and 13:00h using a LI-6400
portable photosynthesis system (LICOR, USA). The LI-6400 chamber was adjusted to a fixed
CO2 concentration (400 µmol mol-1), leaf temperature (31°C) and H2O vapor fraction (21
mmol mol-1), consistent with optimal conditions for photosynthesis in tropical trees (Araújo
et al., 2002; Santos Junior et al., 2006; Doughty & Goulden, 2008; Lloyd & Farquhar, 2008;
Domingues et al., 2014). Asat, E and gs were measured at a photosynthetic photon flux density
(PPFD) of 2000 µmol m-2 s-1 and dark respiration (Rd) at a PPFD of 0 µmol m-2 s-1. Gas
exchange rates stabilized after five minutes of maintaining leaves in the saturated light or
dark inside the IRGA chamber, but we left leaves in the chamber for 30 minutes. While we
could use a model to adjust our measurements to the 25°C convention, we prefer to report the
standard temperature that was in fact used. All measurements were made at a leaf
Accepted Article
This article is protected by copyright. All rights reserved.
temperature of 31°C, which is close to optimal for net photosynthesis (Tribuzy, 2005; Lloyd
& Farquhar, 2008; Slot & Winter, 2017; Tan et al., 2017). We report both photosynthesis and
dark respiration at this leaf temperature. High PPFD guaranteed steady state saturated
photosynthesis in both canopy leaves and in understory trees. We did not detect evidence for
damage to the leaves from this PPFD, i.e. no reduction in photosynthesis with increase up to
2000 µmol m-2 s-1 (Fig. S8). Water use efficiency (WUE) and intrinsic water use efficiency
(WUEi) were calculated as the ratio of light-saturated photosynthesis to transpiration and to
stomatal conductance, respectively. The maximum carboxylation rate of ribulose-1,5-
bisphosphate carboxylase/oxygenase (Vcmax) was estimated for 17 upper canopy trees during
the four 2015 seasons according to the Farquhar et al. model (Farquhar et al., 1980) and the
Sharkey et al. (2007) A/Ci curve fitting utility, version 1.0. For Vcmax, the LI-6400 chamber
was adjusted to a photosynthetic photon flux density of 1000 µmol m-2 s-1, leaf temperature
of 31°C, H2O vapor fraction of 21 mmol mol-1 and successive CO2 concentrations of 400,
300, 250, 200, 150, 100, 50, 400, 400, 450, 500, 600, 700, 800, 1000, 1200 µmol mol-1.
Polyphasic transient of chlorophyll a fluorescence was recorded in three leaves per
tree between 08:00h and 10:00h. Leaves were dark adapted for 30 min, then exposed to a
saturated light pulse of 3000 µmol m-2 s-1 with a wavelength of 650 nm during 1 s using a
portable fluorimeter (PEA, MK2 9600 Hansatech, Norfolk, UK). Chlorophyll a
fluorescence parameters (Maximum quantum yield of photosystem II - Fv/Fm, Performance
index - PIABS, and Total performance index - PItotal) were calculated by the JIP-test (Strasser
et al., 1995; Strasser et al., 2010; Santos Junior et al., 2015). For detailed information about
the fluorescence parameters used in this paper see Strasser et al. (1995); Strasser et al.
(2010); Stirbet & Govindjee (2011) and Stirbet et al. (2018). In summary, during the first
second of exposure to light of a dark-acclimated leaf sample the chlorophyll fluorescence
rises in a polyphasic shape. Four phases or steps of this transient fluorescence signal (O-J-I-
Accepted Article
This article is protected by copyright. All rights reserved.
P) are used to derive the parameter Fv/Fm and the performance indices (PIABS and PItotal). The
Fv/Fm ratio is the maximum quantum yield of the primary Photosystem II photochemistry.
The ratio is derived from the minimum (Fo or O) and the maximum (Fm or P) fluorescence
(Fm-Fo/Fm). The two performance indices are more sensitive but also more complex, being
the product of three parameters from the polyphasic transient curve that define PIABS, plus a
fourth parameter to define PItotal. The PIABS is formed by the ratio of the total number of
active PSII reaction centers (RC) per absorption flux (ABS), Fv/Fm and the probability that an
electron moves further than reduced quinone A [(Fm FJ)/FV)]. Finally, PItotal is calculated
multiplying PIABS by the probability with which an electron from the intersystem electron
carriers is transferred to reduce end electron acceptors at the PSI acceptor side [(Fm FI)/ (Fm
FJ)]. Therefore, PIABS is the potential for energy conservation from photons absorbed by
PSII to the reduction of intersystem electron acceptors and PItotal is the potential for energy
conservation from photons absorbed by PSII to the reduction of PSI end acceptors.
Chloroplast pigment concentrations were determined using a spectrophotometric
method modified from Lichtenthaler & Wellburn (1983), following acetone-filtered
extraction. We used Hendry and Price (1993) equations for chlorophyll (a,b) concentrations.
Leaf nitrogen was determined by the Kjeldahl method with distillation and titration
   ate
method (Murphy & Riley, 1962) and K was determined by atomic absorption spectrometry
(1100B; PerkinElmer, Ueberlingen, Germany).
Additionally, predawn (5:30h - 6:30h) and midday (12:00h 13:00h) leaf water
potentials were measured in 11 upper canopy trees using a pressure pump 3005-1422 (Soil
Moisture Equipment Corp, USA) (Scholander et al., 1964; Turner, 1981) during four months
(May, July, September and October) of the abnormally dry year of 2015. We measured three
Accepted Article
This article is protected by copyright. All rights reserved.
leaves per tree and the canopy walkway allowed each measurement to be completed in about
three minutes.
Statistical analyses
Leaf trait data were subjected to Lilliefors and Mauchly tests to check assumptions of
normality and sphericity, respectively. Seasonal effects were analyzed by one-way repeated
measures ANOVA in each forest layer (upper canopy and understory) and then the seasons
(three 2015 non-drought seasons vs 2015 extreme dry season) were compared by planned
comparison with univariate test of significance (Zar, 1999). Being a repeated-measures test
we included only trees sampled in all seasons for each type of test. We obtained the
difference in photosynthesis between the 2015 dry season and the 2015 annual average
excluding that dry season ( Dry/ Annual)-1) and correlated this difference with leaf traits.
Photosynthesis and stomatal conductance during dry seasons of the 2015 El Niño and the
2016 non-El Niño years were compared by a two-tailed paired sample t-test. All analyses
were performed using Statistica 9.0 software (StatSoft Inc., 2010 East 14th Street, Tulsa, OK,
USA).
Results
Environmental conditions
Starting in August 2015, the El Niño altered the normal Central Amazon rainfall and
air temperature patterns (Fig. 1a, c). However, only the September 2015 field campaign was
preceded by strong drought (Figs. S6, S7). The drought indices ADWD and REW both show
intense water deficits prior to and during the 2015 dry season. ADWD reached -117 mm at
the mid-date of the 2015 dry season field campaign and REW fell to zero. Before and during
the following campaign (dry-to-wet transition), soil water was not yet fully replenished, but
Accepted Article
This article is protected by copyright. All rights reserved.
REW exceeded 0.4 and ADWD oscillated near the  deficit (Fig. S6). Additionally,
predawn leaf water potential -- an indicator of plant-available soil water -- reached its most
negative values during dry months of September and October in 2015 (Fig. S9),
corroborating the two drought indices. In August, September and October of 2015, maximum
monthly air temperatures reached their highest values for the entire 19 y period of record
(Fig. 1c). Rainfall and maximum air temperature were correlated with the Multivariate ENSO
index in a negative and positive way, respectively.
Among the four field campaigns of 2015, the dry season (September) had the highest
mid-day extremes of irradiance (in both upper canopy and understory), highest air
temperature, highest vapor pressure deficit and lowest relative humidity (Fig. 2). Soil
moisture was also lowest in the 2015 dry season at all depths from 10-100 cm (Fig. 2f). In
contrast, the 2016 dry season was not extreme for air temperature, relative humidity or vapor
pressure deficit, having values similar to the dry-to-wet transition campaign of 2015 (Fig. 2c,
d, e). Soil moisture below the surface in the 2016 dry season was similar to that of the pre-
drought wet season of 2015 (Fig. 2f).
Gas exchange
The light-saturated photosynthetic rate varied between the 2015 seasons, reaching a
minimum during the extreme drought of September (Fig. 3a). In the upper canopy and
understory, 93% and 79% of the trees, respectively, reduced their Asat values in this extreme
dry season relative to the mean of the other three seasons in 2015. Across all trees, the dry
season decrease in photosynthesis relative to the other season average was 28% for upper
canopy (F = 10.59; P < 0.01; Fig. 3a) and 17% for understory (F = 11.54; P < 0.01; Fig. 3a).
Two months later, during the dry/wet transition season, in both upper canopy (F = 0.94; P =
0.35) and understory (F = 0.01; P = 0.9), photosynthesis recovered to the pre-drought mean
Accepted Article
This article is protected by copyright. All rights reserved.
value. During the normal dry season of 2016 a non-El Niño year Asat and stomatal
conductance were much higher in the upper canopy and understory when compared with the
2015 El Niño dry season (Fig 4).
Following the trends in Asat, a high percentage of trees reduced their stomatal
conductance (gs) during the dry season. Seventy-three percent of upper canopy trees reduced
gs in the wet-to-dry transition season and 88% kept low values through the extreme dry
season of 2015. In the understory, a high percentage of trees with reduced gs (83%) were seen
only during the extreme dry season. Means of gs and of E across all understory trees declined
in the extreme dry season of 2015 relative to the three other seasons (gs F = 32.89; P <
0.0001; Fig. 3b; E F = 23.31; P < 0.0001; Fig. 3d). In the upper canopy, however, these two
variables were lower in both the wet-to-dry transition season and the extreme dry season,
when compared to the other season average (gs F = 20.86; P < 0.001; E F = 26.59; P < 0.001).
In the post-drought dry-to-wet transition season, gs recovered to wet season values in both
forest layers (Understory - F = 2.93; P = 0.10; Upper Canopy - F = 4.56; P = 0.05). However,
E recovered to wet season values only in the understory (Understory - F = 0.82; P = 0.37;
Canopy - F = 5.17; P = 0.04). For upper canopy trees, the WUE and WUEi increased in the
wet-to-dry transition season and maintained high efficiency through the extreme dry and dry-
to-wet seasons of 2015 (Fig. 3e,f). In understory, the trees increased their WUE and WUEi
only during the extreme dry season, but the magnitude of increase was similar to upper
canopy trees (Fig. 3e,f).
Rd remained unchanged through all four seasons of 2015 in both upper canopy and
understory (Fig. 3c). Similarly, no significant seasonal changes in Vcmax were observed in
upper canopy trees species (Fig. S4).
Accepted Article
This article is protected by copyright. All rights reserved.
Chlorophyll concentration and fluorescence
The maximum quantum yield of photosystem II (Fv/Fm) ranged from 0.73 to 0.85 and
the upper canopy mean quantum yield did not differ among the four 2015 seasons (Fig. 5b).
On the other hand, mean Fv/Fm in the understory was different between the year 2015 dry-to-
wet and the extreme dry season (F = 35.79; P < 0.001). PIABS, PItotal and leaf chlorophyll
concentration did not change with the seasons (Fig. 5a, c, d).
Leaf nutrients concentration
In both upper canopy and understory leaves, mean leaf N concentration did not reduce
during the 2015 extreme dry season (Fig. 6a). P concentration was lowered during wet and
dry-to-wet seasons, but only in the understory (F = 6.39; P < 0.01; Fig. 6b). In the understory,
during the extreme 2015 dry season, mean leaf K concentration was slightly higher compared
with all other seasons (F = 2.58; P = 0.08; Fig. 6c).
Relationships between reduction in Asat and leaf traits
Asat reduction during the 2015 dry season was correlated only with a reduced stomatal
conductance. This was true in both upper canopy and understory (Table S3; Fig. 7). Seasonal
changes in photochemical performance parameters (Fv/Fm, PIABS, PItotal), chlorophyll
concentration and leaf nutrient content were not correlated with Asat reduction (Table S3).
Discussion
El Niño drought inhibits leaf level photosynthesis in Central Amazon trees
We have shown that a strong El Niño drought in the dry season of 2015 inhibited leaf
level photosynthesis (Asat) in mature leaves of a Central Amazon forest. This was not a
normal seasonal pattern; Asat reduction did not occur during the climatically normal dry
Accepted Article
This article is protected by copyright. All rights reserved.
season of the following year. Extreme drought and/or concomitant high air temperature
induced a diffusive restriction to CO2 assimilation by forcing stomatal closure (Fig. 2 and
Fig. 3). This is consistent with previous reports of photosynthesis reduction during extreme
droughts (Stahl et al., 2013a; Gatti et al., 2014, Aragão et al., 2014; Hilker et al., 2014;
Doughty et al., 2015; Inoue et al., 2017) or warming (Doughty, 2011; Cavaleri et al., 2015).
We found no evidence that photosynthesis reduction could be attributed to causes beyond
stomatal closure, i.e., declines in photosynthetic infrastructure such as reduced photosynthetic
capacity, photochemical performance, chlorophyll concentration or leaf nutrient content
(Table S2).
This is an important finding: even during        
(Fig. 1; Jiménez-Muñoz et al., 2016), the only detectable mechanism of photosynthetic
reduction was stomatal closure, a transient regulatory response that can potentially reverse
when the drought ends.
Stomatal conductance as the main driver of photosynthesis reduction during drought
Stomatal closure-induced reductions in leaf level photosynthesis were sensitive to
canopy layer (Fig. 3a). Reduced photosynthesis, stomatal conductance and transpiration were
more pronounced in upper canopy than in understory trees (Fig. 3c and Fig. 3d). Both upper
canopy and understory trees could have deep roots allowing access to deep soil water stores
(Nepstad et al., 1994, Stahl et al., 2013b). However, the typical microclimate conditions in
the upper canopy layer are low air humidity, high air temperature and high irradiance
(Chazdon & Fetcher, 1984; Kumagai et al., 2001; Kamakura et al., 2015), all of which are
aggravated during the dry season, especially in an El Niño year (Fig. 2). Thus, the more upper
canopy leaves are exposed to demanding atmospheric conditions, the more likely they are to
exhibit a strong reduction of stomatal conductance and photosynthesis during drought,
Accepted Article
This article is protected by copyright. All rights reserved.
compared to understory trees (Sperlich et al., 2015). Moreover, taller trees must cope with a
longer water-transport path from the soil to their leaves. Higher atmospheric evaporative
demand increases the risk of embolism and hydraulic failure (Bennett et al., 2015). In sum,
our results provide support for more pronounced drought effects on leaf gas exchange in tall
upper canopy trees, exposed to drier and hotter atmosphere, than in understory trees.
Our focus on individual mature leaves means that several important attributes of
whole canopy photosynthesis were outside the scope of this study. We do not here examine,
for example, whether drought influences canopy phenology (leaf growth, mortality, or leaf
age-composition). In a normal (non-El Niño) year, dry season leaf flush drives a dry season
increase in whole-canopy photosynthesis at this site (Restrepo-Coupe et al., 2013; Wu et al.,
2016). Newly flushed leaves -- if supplied sufficient water during this 2015 El Niño dry
season (as in more typical dry seasons) -- would counteract some of the leaf-level suppression
of photosynthetic Asat, potentially allowing canopy photosynthesis to increase despite a
drought.
An important result of our study is that it shows, for the first time, that photosynthetic
reduction at the leaf-level during an El Niño drought was caused exclusively by stomatal
closure. However, the relative importance of drought and of high temperature on
photosynthesis of Amazon trees are still not fully elucidated. In parts of the Eastern Amazon
normal seasonal dry periods last five months but do not limit photosynthesis (Domingues et
al., 2014). In a 15y long through-fall exclusion experiment, where soil water was artificially
reduced but upper canopy air temperature was unaffected, only vulnerable trees reduced
photosynthesis in association with a drop in stomatal conductance (Rowland et al., 2015a).
Trees died by embolism without any pre-death reduction of growth rate or of non-structural
carbohydrate stores (Rowland et al., 2015b). By contrast, during a single year of extreme
drought in 2010 in western Amazonia, photosynthesis was suppressed at leaf and plot levels
Accepted Article
This article is protected by copyright. All rights reserved.
with an estimated assimilation reduction of 0.38 petagrams of carbon across the Amazon
Basin (Doughty et al., 2015).
Non-stomatal processes were not impaired during extreme drought in Central Amazon trees
The cause of stomatal closure during the dry season was presumably to maintain leaf
water potential under low available soil water and high vapor pressure deficit, as observed in
11 canopy trees in this survey and previously reported for tropical tree species (Fig S9; Reich
& Borchert 1988; Bonal et al., 2000; Cao, 2000; Miranda et al., 2005). With increasing
severity and duration of the drought, other processes related to photosynthesis can be
affected, such as photochemistry in the thylakoid membranes (Flexas & Medrano 2002;
Flexas et al., 2006; Atkin & Macherel 2009). For example, in evergreen and semi-deciduous
trees of a monsoonal tropical dry forest the electron transport rate (ETR) decreased in
response to drought stress during the dry season (Ishida et al., 2014). Further, Hung et al.
(2013) demonstrated an inhibition of both linear electron flow and non-photochemical
quenching under prolonged drought for tropical trees growing on limestone. However,
according to JIP-test parameters, the electron transport performance was not impaired during
the extreme 2015 dry season (Fig. 5a, b, d). The JIP-test has been used in several studies that
analyzed the effects of drought stress on the integrity and functionality of electron transport
(Strasser et al., 2010; Redillas et al., 2011; Desotgiu et al., 2012; Campos et al., 2014). PItotal
integrates the energy conservation from photons absorbed by PSII through to the reductions
of PSI end acceptors and is among the more sensitive JIP-test parameters to biotic and abiotic
stresses (Strasser et al., 2010; Stirbet & Govindjee 2011; Stirbet et al., 2018). The uncoupling
between electron transport and photosynthesis during the dry season suggests that another
process, such as photorespiration, was contributing to energy and power reduction (ATP,
NADPH, Fd-) utilization generated in the electron transport chain (Martinez-Ferii et al., 2000,
Accepted Article
This article is protected by copyright. All rights reserved.
Haupt-Herting & Fock, 2002; Lawlor & Tezara, 2009). Moreover, this uncoupling can add
more complexity for obtaining landscape-level photosynthesis estimates based on the
chlorophyll fluorescence signal (Osuna et al., 2015; Yang et al., 2017).
In conclusion, our results have important implications for improving models of the
sensitivity of tropical forests to climate anomalies by providing insight into the effects of
extreme climatic events (e.g. El Niño) on leaf-level photosynthesis. This work also adds some
early evidence about the consequences of future climate change on the current carbon cycle
of Amazonia. Because stomatal conductance best explained changes in photosynthesis as
affected by drought stress, including this variable in models may lead to improved model
performance under drought stress conditions for Amazon forest.
Acknowledgments
We thank the Federal University of Amazonas (UFAM), Brazil´s National Institute
for Amazon Research (INPA) and the Large-Scale Biosphere-Atmosphere Experiment in
Amazonia Program (LBA) for logistic support; the GOAmazon project, funded jointly by the
U.S. Department of Energy (DOE, # DE. SC0008383), by the Fundação de Amparo à
Pesquisa do Estado de São Paulo (FAPESP), and by the Fundação de Amparo à Pesquisa do
Estado do Amazonas (FAPEAM, # 062.00570/2014). The authors declare no conflict of
interest.
References
     
Gloor, E. (2014). Environmental change and the carbon balance of Amazonian forests.
Biological Reviews, 89, 913931. https://doi.org/10.1111/brv.12088
Accepted Article
This article is protected by copyright. All rights reserved.
Araújo, A. C.
(2002). Comparative measurements of carbon dioxide fluxes from two nearby towers in a
central Amazonian rainforest: The Manaus LBA site. Journal of Geophysical Research,
107(D20), 120. https://doi.org/10.1029/2001JD000676
Atkin, O. K., & Macherel, D. (2008). The crucial role of plant mitochondria in orchestrating
drought tolerance. Annals of botany, 103, 581-597. https://doi.org/10.1093/aob/mcn094
Bennett, A. C., McDowell, N. G., Allen, C. D., & Anderson-Teixeira, K. J. (2015). Larger
trees suffer most during drought in forests worldwide. Nature Plants, 1(10), 15139.
https://doi.org/10.1038/nplants.2015.139
Bonal, D., Barigah, T. S., Granier, A., & Guehl, J. M. (2000). Late-stage canopy tree species
al
rainforest of French Guiana. Plant, Cell and Environment, 23, 445459.
https://doi.org/10.1046/j.1365-3040.2000.00556.x
Bonal, D., Bosc, A., Ponton, S., Goret, J. Y., Burban, B., Gross, P., Granier, A. (2008).
Impact of severe dry season on net ecosystem exchange in the Neotropical rainforest of
French Guiana. Global Change Biology, 14, 19171933. http://dx.doi.org/10.1111/j.1365-
2486.2008.01610.x
Bonal, D., Burban, B., Stahl, C., Wagner, F., & Hérault, B. (2016). The response of tropical
rainforests to droughtlessons from recent research and future prospects. Annals of Forest
Science, 73(1), 2744. https://doi.org/10.1007/s13595-015-0522-5
Bremner, J. M. (1996). Nitrogen-Total. In D.L. Sparks (Ed.), Methods of soil analysis. Part 3.
Chemical Methods. (pp. 1085-1121).
Brienen, R. J. W., Phillips, O. L., Feldpausch, T. R., Gloor, E., Baker, T. R., Lloy  
Zagt, R. J. (2015). Long-term decline of the Amazon carbon sink. Nature, 519, 344348.
https://doi.org/10.1038/nature14283
Accepted Article
This article is protected by copyright. All rights reserved.
Campos, H., Trejo, C., Peña-Valdivia, C. B., García-Nava, R., Conde-Martínez, F. V., &
Cruz-Ortega, M. R. (2014). Stomatal and non-stomatal limitations of bell pepper (Capsicum
annuum L.) plants under water stress and re-watering: Delayed restoration of photosynthesis
during recovery. Environmental and Experimental Botany, 98, 5664.
https://doi.org/10.1016/j.envexpbot.2013.10.015
Cao, K. (2000). Water relations and gas exchange of tropical saplings during a prolonged
drought in a Bornean heath forest, with reference to root architecture. Journal of Tropical
Ecology, 16, 101116. https://doi.org/10.1017/S0266467400001292
Cavaleri, M. A., Coble, A. P., Ryan, M. G., Bauerle, W. L., Loescher, H. W. & Oberbauer, S.
F. (2017), Tropical rainforest carbon sink declines during El Niño as a result of reduced
photosynthesis and increased respiration rates. New Phytologist, 216, 136149.
http://dx.doi.org/10.1111/nph.14724
Cavaleri, M. A., Reed, S. C., Smith, W. K. & Wood, T. E. (2015). Urgent need for warming
experiments in tropical forests. Global Change Biology, 21, 21112121.
https://doi.org/10.1111/gcb.12860
Chambers, J. Q., Trib
Trumbore, S. E. (2004). Respiration from a tropical forest ecosystem: Partitioning of sources
and low carbon use efficiency. Ecological Applications, 14, 72-88.
https://doi.org/10.1890/01-6012
Chastain, D. R., Snider, J. L., Collins, G. D., Perry, C. D., Whitaker, J., & Byrd, S. A. (2014).
Water deficit in field-grown Gossypium hirsutum primarily limits net photosynthesis by
decreasing stomatal conductance, increasing photorespiration, and increasing the ratio of dark
respiration to gross photosynthesis. Journal of Plant Physiology, 171, 15761585.
https://doi.org/10.1016/j.jplph.2014.07.014
Accepted Article
This article is protected by copyright. All rights reserved.
Chazdon, R., & Fetcher, N. (1984). Photosynthetic light environments in a lowland tropical
rain forest in Costa Rica. Journal of Ecology, 72, 553564. Retrieved from
http://www.jstor.org/discover/10.2307/2260066?uid=2&uid=4&sid=21104355119067
Corlett, R. T. (2016). The Impacts of Droughts in Tropical Forests. Trends in Plant Science,
21, 584593. https://doi.org/10.1016/j.tplants.2016.02.003
Davidson, E. A., de Araújo, A. C., Artaxo, P., Balch, J. K., Brown, I. F., C. Bustamante, M.
  Wofsy, S. C. (2012). The Amazon basin in transition. Nature, 481, 321328.
https://doi.org/10.1038/nature10717
Desotgiu, R., Pollastrini, M., Cascio, C., Gerosa, G., Marzuoli, R., & Bussotti, F. (2012).
Chlorophyll a fluorescence analysis along a vertical gradient of the crown in a poplar (Oxford
clone) subjected to ozone and water stress. Tree Physiology, 32, 976986.
https://doi.org/10.1093/treephys/tps062
Domingues, T. F., Martinelli, L. A., & Ehleringer, J. R. (2014). Seasonal patterns of leaf-
level photosynthetic gas exchange in an eastern Amazonian rain forest. Plant Ecology &
Diversity, 7, 189203. https://doi.org/10.1080/17550874.2012.748849
Doughty, C. E., & Goulden, M. L. (2008). Seasonal patterns of tropical forest leaf area index
and CO2 exchange. Journal of Geophysical Research, 113, G00B06.
https://doi.org/10.1029/2007JG000590
Doughty, C. E. (2011). An In Situ Leaf and Branch Warming Experiment in the Amazon.
Biotropica, 43: 658665. https://doi.org/10.1111/j.1744-7429.2010.00746.x
Doughty, C. E., Metcalfe, D. B., Girardin, C. A. J., Amézquita, F. F., Cabrera, D. G., Huasco,
   Malhi, Y. (2015). Drought impact on forest carbon dynamics and fluxes in
Amazonia. Nature, 519, 7882. https://doi.org/10.1038/nature14213
Accepted Article
This article is protected by copyright. All rights reserved.
Farquhar, G. D., Von Caemmerer, S., & Berry, J. A. (1980). A Biochemical Model of
Photosynthetic CO2 Assimilation in Leaves of C3 Species. Planta, 149, 7890.
https://doi.org/10.1007/BF00386231
Fauset, S., Johnson, M. O., Gloor, M., Baker, T. R., Monteagudo M, A., 
Phillips, O. L. (2015). Hyperdominance in Amazonian forest carbon cycling. Nature
Communications, 6, 6857. https://doi.org/10.1038/ncomms7857
Flexas, J., & Medrano, H. (2002). Drought-inhibition of photosynthesis in C3 plants:
Stomatal and non-stomatal limitations revisited. Annals of Botany, 89, 183189.
https://doi.org/10.1093/aob/mcf027
Flexas, J., Bota, J., Galmés, J., Medrano, H., & Ribas-Carbó, M. (2006). Keeping a positive
carbon balance under adverse conditions: Responses of photosynthesis and respiration to
water stress. Physiologia Plantarum, 127, 343-352. https://doi.org/10.1111/j.1399-
3054.2006.00621.x

J. (2014). Drought sensitivity of Amazonian carbon balance revealed by atmospheric
measurements. Nature, 506, 7680. https://doi.org/10.1038/nature1295
Granier, A., Bréda, N., Biron, P., & Villette, S. (1999). A lumped water balance model to
evaluate duration and intensity of drought constraints in forest stands. Ecological Modelling,
116: 269-283. https://doi.org/10.1016/S0304-3800(98)00205-1
Haupt-Herting, S., & Fock, H. P. (2002). Oxygen exchange in relation to carbon assimilation
in water-stressed leaves during photosynthesis. Annals of Botany, 89, 851859.
https://doi.org/10.1093/aob/mcf023
Hendry, G. A. F., & Price, A. H. (1993). Stress indicators: chlorophylls and carotenoids. In
G. A. F. Hendry & J.P. Grime (Eds.), Methods in Comparative Plant Ecology (pp. 148-152).
Accepted Article
This article is protected by copyright. All rights reserved.
Hilker, 
J. (2014). Vegetation dynamics and rainfall sensitivity of the Amazon. Proceedings of the
National Academy of Sciences, 111, 1604116046. https://doi.org/10.1073/pnas.1404870111
Huang, W., Fu, P. L., Jiang, Y. J., Zhang, J. L., Zhang, S. B., Hu, H., & Cao, K. F. (2013).
Differences in the responses of photosystem I and photosystem II of three tree species
Cleistanthus sumatranus, Celtis philippensis and Pistacia weinmannifolia exposed to a
prolonged drought in a tropical limestone forest. Tree physiology, 33, 211-220.
https://doi.org/10.1093/treephys/tps132
Inoue, Y., Ichie, T., Kenzo, T., Yoneyama, A., Kumagai, T. O., & Nakashizuka, T. (2017).
Effects of rainfall exclusion on leaf gas exchange traits and osmotic adjustment in mature
canopy trees of Dryobalanops aromatica (Dipterocarpaceae) in a Malaysian tropical rain
forest. Tree physiology, 37, 1301-1311. https://doi.org/10.1093/treephys/tpx053
Ishida, A., Yamazaki, J. Y., Harayama, H., Yazaki, K., Ladpala, P., Nakano, T., ... Maeda, T.
(2013). Photoprotection of evergreen and drought-deciduous tree leaves to overcome the dry
season in monsoonal tropical dry forests in Thailand. Tree physiology, 34, 15-28.
https://doi.org/10.1093/treephys/tpt107
Jiménez-Muñoz, J. C., Mattar, C., Barichivich, J., Santamaría-Artigas, A., Takahashi, K.,
Schrier, G. van der. (2016). Record-breaking warming and extreme drought in
the Amazon rainforest during the course of El Niño 20152016. Scientific Reports, 6 (33130),
1-7. https://doi.org/10.1038/srep33130
Kamakura, M., Kosugi, Y., Takanashi, S., Uemura, A., Utsugi, H., & Kassim, A. R. (2015).
Occurrence of stomatal patchiness and its spatial scale in leaves from various sizes of trees
distributed in a South-east Asian tropical rainforest in Peninsular Malaysia. Tree Physiology,
35, 6170. https://doi.org/10.1093/treephys/tpu109
Accepted Article
This article is protected by copyright. All rights reserved.
Kitajima, K., Mulkey, S. S., & Wright, S. J. (1997). Decline of photosynthetic capacity with
leaf age in relation to leaf longevities for five tropical canopy tree species. American Journal
of Botany, 84, 702708. https://doi.org/10.2307/2445906
Kumagai, T., Kuraji, K., Noguchi, H., Tanaka, Y., Tanaka, K., & Suzuki, M. (2001). Vertical
profiles of environmental factors within tropical rainforest, Lambir Hills National Park,
Sarawak, Malaysia. Journal of Forest Research, 6, 257264.
https://doi.org/10.1007/BF02762466
Lawlor, D. W., & Tezara, W. (2009). Causes of decreased photosynthetic rate and metabolic
capacity in water-deficient leaf cells: A critical evaluation of mechanisms and integration of
processes. Annals of Botany, 103, 561-579. https://doi.org/10.1093/aob/mcn244
Lewis, S. L., Brando, P. M., Phillips, O. L., van der Heijden, G. M., & Nepstad, D. (2011).
The 2010 amazon drought. Science, 331(6017), 554-554.
https://doi.org/10.1126/science.1200807
Lichtenthaler, H., & Wellburn, A. (1983). Determinations of total carotenoids and
chlorophylls b of leaf extracts in different solvents. Biochemical Society Transactions, 11,
591592. https://doi.org/10.1042/bst0110591
Lloyd, J., & Farquhar, G. D. (2008). Effects of rising temperatures and [CO2] on the
physiology of tropical forest trees. Philosophical Transactions of the Royal Society of
London. Series B, Biological Sciences, 363, 18111817.
https://doi.org/10.1098/rstb.2007.0032
                  
Kruijt, B. (2004). Variation of carbon and nitrogen cycling processes along a topographic
gradient in a central Amazonian forest. Global Change Biology, 10, 592600.
https://doi.org/10.1111/j.1529-8817.2003.00757.x
Accepted Article
This article is protected by copyright. All rights reserved.
Malhi, Y., Aragão, L. E., Galbraith, D., Huntingford, C., Fisher, R., Zelazowski, P., ... &
Meir, P. (2009). Exploring the likelihood and mechanism of a climate-change-induced
dieback of the Amazon rainforest. Proceedings of the National Academy of Sciences, 106,
20610-20615. https://doi.org/10.1073/pnas.0804619106
Malhi, Y., Roberts, J. T., Betts, R. A., Killeen, T. J., Li, W., & Nobre, C. A. (2008). Climate
change, deforestation, and the fate of the Amazon. Science, 319, 169172.
https://doi.org/10.1126/science.1146961
Malhi, Y., Wood, D., Baker  
(2006). The regional variation of aboveground live biomass in old-growth Amazonian
forests. Global Change Biology, 12, 11071138. https://doi.org/10.1111/j.1365-
2486.2006.01120.x
Martínez-Ferri, E., Balaguer, L., Valladares, F., Chico, J. M., & Manrique, E. (2000). Energy
dissipation in drought-avoiding and drought-tolerant tree species at midday during the
Mediterranean summer. Tree Physiology, 20, 131138.
https://doi.org/10.1093/treephys/20.2.131
     
Shiraiwa, S. (2005). Seasonal variation in the leaf gas exchange of tropical forest trees in the
rain forestsavanna transition of the southern Amazon Basin. Journal of Tropical Ecology,
21, 451460. https://doi.org/10.1017/S0266467405002427
Murphy, J., & Riley, J. P. (1962). A modified single solution method for the determination of
phosphate in natural waters. Analytica Chimica Acta, 27, 3136.
https://doi.org/10.1016/S0003-2670(00)88444-5
NASA (2016) Tropical Rainfall Measuring Mission Project (TRMM), 3B42v7 and 3B43v7.
NASA Distrib. Active Arch. Cent., Goddard Space Flight Cent. Earth Sci., Greenbelt, Md.
Accepted Article
This article is protected by copyright. All rights reserved.
Nepstad, D. C., de Carvalho, C. R., Davidson, E. A., Jipp, P. H., Lefebvre, P. A., Negreiros,
Vieira, S. (1994). The role of deep roots in the hydrological and carbon cycles of
Amazonian forests and pastures. Nature, 372, 666669. https://doi.org/10.1038/372666a0
Osuna, J. L., Baldocchi, D. D., Kobayashi, H., & Dawson, T. E. (2015). Seasonal trends in
photosynthesis and electron transport during the Mediterranean summer drought in leaves of
deciduous oaks. Tree Physiology, 35, 485500. https://doi.org/10.1093/treephys/tpv023
Oukarroum, A., Schansker, G., & Strasser, R. J. (2009). Drought stress effects on
photosystem I content and photosystem II thermotolerance analyzed using Chl a fluorescence
kinetics in barley varieties differing in their drought tolerance. Physiologia Plantarum, 137,
188189. https://doi.org/10.1111/j.1399-3054.2009.01273.x
Phillips, O. L., Aragão, L. E. O. C., Lewis, S. L., Fisher, J. B., Lloyd, J., Lopez-Gonzalez, G.,
-Lezama, A. (2009). Drought Sensitivity of the Amazon Rainforest. Science, 323,
13441347. https://doi.org/10.1126/science.1164033
Redillas, M. C. F. R., Strasser, R. J., Jeong, J. S., Kim, Y. S., & Kim, J. K. (2011). The use of
JIP test to evaluate drought-tolerance of transgenic rice overexpressing OsNAC10. Plant
Biotechnology Reports, 5, 169175. https://doi.org/10.1007/s11816-011-0170-7
Reich, P. B., & Borchert, R. (1988). Changes with leaf age in stomatal function and water
status of several tropical tree species. Biotropica, 20, 6069.
https://doi.org/10.2307/23884277
Restrepo-Coupe, N., da Rocha, H. R., Hutyra, L. R., da Araujo, A. C., Borma, L. S.,
   Saleska, S. R. (2013). What drives the seasonality of photosynthesis
across the Amazon basin? A cross-site analysis of eddy flux tower measurements from the
Brasil flux network. Agricultural and Forest Meteorology, 182183, 128144.
https://doi.org/10.1016/j.agrformet.2013.04.031
Accepted Article
This article is protected by copyright. All rights reserved.
Rowland, L., da Costa, A. C. L., Galbraith, D. R., Oliveira, R. S., Binks, O. J., Oliveira, A. A.
Meir, P. (2015b). Death from drought in tropical forests is triggered by hydraulics not
carbon starvation. Nature, 528, 119122. https://doi.org/10.1038/nature15539
Rowland, L., Lobo-do-Vale, R. L., Christoffersen, B. O., Melém, E. A., Kruijt, B.,
Meir, P. (2015a). After more than a decade of soil moisture deficit,
tropical rainforest trees maintain photosynthetic capacity, despite increased leaf respiration.
Global Change Biology, 21, 46624672. https://doi.org/10.1111/gcb.13035
Saleska, S. R., Didan, K., Huete, A. R., & da Rocha, H. R. (2007). Amazon Forests Green-Up
During 2005 Drought. Science, 318, 618. https://doi.org/10.1126/science.1146663
Santos Junior, U. M., Gonçalves, J. F. D. C., Strasser, R. J., & Fearnside, P. M. (2015).
Flooding of tropical forests in central Amazonia: what do the effects on the photosynthetic
apparatus of trees tell us about species suitability for reforestation in extreme environments
created by hydroelectric dams? Acta Physiologiae Plantarum, 37, 166.
https://doi.org/10.1007/s11738-015-1915-7
Santos, U. M., de Carvalho Gonçalves, J. F., & Feldpausch, T. R. (2006). Growth, leaf
nutrient concentration and photosynthetic nutrient use efficiency in tropical tree species
planted in degraded areas in central Amazonia. Forest Ecology and Management, 226, 299
309. https://doi.org/10.1016/j.foreco.2006.01.042
Scholander, P. F., Hammel, H. T., Bradstreet, D., Hemmingsen, E. A. (1964). Sap Pressure in
Vascular Plants Negative hydrostatic pressure can be measured in plants. Science 148: 339
346. https://doi.org/10.1126/science.148.3668.339
Sharkey, T. D., Bernacchi, C. J., Farquhar, G. D., & Singsaas, E. L. (2007). Fitting
photosynthetic carbon dioxide response curves for C3 leaves. Plant, Cell and Environment,
30, 10351040. https://doi.org/10.1111/j.1365-3040.2007.01710.x
Accepted Article
This article is protected by copyright. All rights reserved.
Slot, M. & Winter, K. (2017). In situ temperature relationships of biochemical and stomatal
controls of photosynthesis in four lowland tropical tree species. Plant, Cell & Environment,
40, 3055-3068. https://doi.org/10.1111/pce.13071
Sperlich, D., Chang, C. T., Peñuelas, J., Gracia, C., & Sabaté, S. (2015). Seasonal variability
of foliar photosynthetic and morphological traits and drought impacts in a Mediterranean
mixed forest. Tree Physiology, 35, 501520. https://doi.org/10.1093/treephys/tpv017
Sperlich, D., Chang, C. T., Peñuelas, J., Gracia, C., & Sabaté, S. (2015). Seasonal variability
of foliar photosynthetic and morphological traits and drought impacts in a Mediterranean
mixed forest. Tree Physiology, 35, 501520. https://doi.org/10.1093/treephys/tpv017
Stahl, C., Burban, B., Wagner, F., Goret, J.-Y., Bompy, F. and Bonal, D. (2013a). Influence
of seasonal variations in soil water availability on gas exchange of tropical canopy trees.
Biotropica, 45, 155164. doi:10.1111/j.1744-7429.2012.00902.x
Stahl, C., Hérault, B., Rossi, V., Burban, B., Bréchet, C., & Bonal, D. (2013b). Depth of soil
water uptake by tropical rainforest trees during dry periods: Does tree dimension matter?
Oecologia, 173, 11911201. https://doi.org/10.1007/s00442-013-2724-6
Stirbet, A., & Govindjee (2011) On the relation between the Kautsky effect (chlorophyll a
fluorescence induction) and photosystem II: basics and applications of the OJIP fluorescence
transient. Journal of Photochemistry and Photobiology B: Biology, 104, 236-257.
https://doi.org/10.1016/j.jphotobiol.2010.12.010
Stirbet, A., Lazár, D., Kromdijk J, & Govindjee (2018) Chlorophyll a fluorescence induction:
Can just a one-second measurement be used to quantify abiotic stress responses?
Photosynthetica. https://doi.org/10.1007/s11099-018-0770-3.
Strasser, R. J., Srivastava, A., & Govindjee. (1995). Polyphasic chlorophyll a fluorescence
transient in plants and cyanobacteria. Photochemistry and Photobiology, 61, 3242.
https://doi.org/10.1111/j.1751-1097.1995.tb09240.x
Accepted Article
This article is protected by copyright. All rights reserved.
Strasser, R. J., Tsimilli-Michael, M., Qiang, S., & Goltsev, V. (2010). Simultaneous in vivo
recording of prompt and delayed fluorescence and 820-nm reflection changes during drying
and after rehydration of the resurrection plant Haberlea rhodopensis. Biochimica et
Biophysica Acta - Bioenergetics, 1797, 13131326.
https://doi.org/10.1016/j.bbabio.2010.03.008
Tan, Z. H., Zeng, J., Zhang, Y. J., Slot, M., Gamo, M., Hirano, T., ... Wofsy, S. C. (2017).
Optimum air temperature for tropical forest photosynthesis: Mechanisms involved and
implications for climate warming. Environmental Research Letters, 12, 054022.
https://doi.org/10.1088/1748-9326/aa6f97
Teskey, R., Wertin, T., Bauweraerts, I., Ameye, M., McGuire, M. A., & Steppe, K. (2015).
Responses of tree species to heat waves and extreme heat events. Plant, Cell and
Environment, 38, 1699-1712. https://doi.org/10.1111/pce.12417
Tribuzy, E. S. (2005). Variações da temperatura foliar do dossel e o seu efeito na taxa
assimilatória de CO2 na Amazônia Central. Doctoral Thesis, Ecologia de Agroecossistemas,
University of São Paulo, Piracicaba. doi:10.11606/T.91.2005.tde-15072005-144011.
Retrieved 2017-11-06, from www.teses.usp.br
Turner, N. C. 1981. Techniques and experimental approaches for the measurement of plant
water status. Plan and Soil, 366: 339366. https://doi.org/10.1007/BF02180062

(2004). Forest structure and carbon dynamics in Amazonian tropical rain forests. Oecologia,
140, 468479. https://doi.org/10.1007/s00442-004-1598-z
Wagner, F., Hérault, B., Stahl, C., Bonal, D., & Rossi, V. (2011). Modeling water availability
for trees in tropical forests. Agricultural and Forest Meteorology, 151: 1202-1213.
https://doi.org/10.1016/j.agrformet.2011.04.012
Accepted Article
This article is protected by copyright. All rights reserved.
Wagner, F. H., Hérault, B., Bonal, D., Stahl, C., Anderson, L. O., Baker, T. R., ... Bowman,
D. M. (2016). Climate seasonality limits leaf carbon assimilation and wood productivity in
tropical forests. Biogeosciences, 13, 2537-2562. https://doi.org/10.5194/bg-13-2537-2016
Wagner, F. H., Hérault, B., Rossi, V., Hilker, T., Maeda, E. E., Sanchez, A, ... Aragão, L. E.
O. C. (2017) Climate drivers of the Amazon forest greening. PLoS ONE 12(7), e0180932.
https://doi.org/10.1371/journal.pone.0180932
Wolter, K., & Timlin, M. S. (1998). Measuring the strength of ENSO events: How does
1997/98 rank? Weather, 53, 315324. https://doi.org/10.1002/j.1477-8696.1998.tb06408.x
Wu, J., Albert, L. P., Lopes, A. P., Restrepo-Coupe, N., Hayek, M., Wiedem 
Saleska, S. R. (2016). Leaf development and demography explain photosynthetic seasonality
in Amazon evergreen forests. Science, 351, 972976. https://doi.org/10.1126/science.aad5068
Wu, J., Albert, L. P., Lopes, A. P., Restrepo-Coupe, N., Hay  
Saleska, S. R. (2016). Leaf development and demography explain photosynthetic seasonality
in Amazon evergreen forests. Science, 351, 972976. https://doi.org/10.1126/science.aad5068
Yang, H., Yang, X., Zhang, Y., Heskel, M. A.,   
Chlorophyll fluorescence tracks seasonal variations of photosynthesis from leaf to canopy in
a temperate forest. Global Change Biology, 23, 28742886.
https://doi.org/10.1111/gcb.13590
Zhao, M., & Running, S. W. (2010). Drought-Induced Reduction in Global. Science, 329,
940943. https://doi.org/10.1126/science.1192666
Accepted Article
This article is protected by copyright. All rights reserved.
Fig. 1 Monthly average rainfall (a) and maximum monthly air temperature (c) in 2015 (bars).
Monthly averages for 19 y (1998-2016) are solid lines; the range for each month´s average
rainfall and for each month´s maximum air temperature are dashed lines. Anomalies for 19 y
of trimester average precipitation (b) and trimester maximum air temperature (d) are plotted
against trimester anomalies for Multivariate ENSO Index MEI. The four 2015 trimesters
are solid black symbols (JFM - inverted triangle; AMJ - triangle; JAS - square and OND -
circle).
Fig. 2 Diurnal variation in photosynthetic photon flux density above the canopy (PPFD; a)
and in the understory (PPFD; b); air temperature (Tair; c), relative humidity (RH; d) and
vapor pressure deficit (VPD; e) above the canopy; and soil moisture (f) of a Central Amazon
forest during the four field seasons of the 2015 El Niño year and dry season of the 2016 non-
El Niño year. Symbols with black outlines at each time of day are the mean for that time of
day for each of the five field campaigns. Light grey symbols are daily values.
Fig. 3 Box plots of light-saturated photosynthetic rate (Asat; a), stomatal conductance (gs; b),
leaf dark respiration (Rd; c); transpiration (E; d), intrinsic water use efficiency (WUEi, e) and
water use efficiency (WUE; f) in upper canopy and understory trees of a Central Amazon
forest during the four field seasons of the 2015 El Niño year. Solid line inside the box is the
median, dashed line is the mean, box-plot range is 50% of the data, error bars are 5th and 95th
percentiles and points are outliers. Asterisk p-values are n.s.(p>0.05), *(p<0.05), **(p<0.01),
***(p<0.001) for a one-way repeated-measures ANOVA, i.e., testing for a difference
between any pair of seasons (Degrees of freedom are shown in Table S2). Comparisons
between dry season and the other three seasons grouped together are reported in the text.
Accepted Article
This article is protected by copyright. All rights reserved.
Fig 4 Box plots of light-saturated photosynthetic rate (Asat; a) and stomatal conductance (gs;
b) during the 2015 (El Niño year) and the 2016 (non-El Niño year) dry seasons for the upper
canopy and understory trees. For box interpretation see Fig 3. The P-values are for two-tailed
paired sample t-tests (nunderstory = 21; ncanopy = 14).
Fig. 5 Box plots of chlorophyll concentration (Chl; a), maximum quantum yield of
photosystem II (Fv/Fm; b), photochemical performance index (PIABS; c), and total
photochemical performance index (PItotal; d) in upper canopy and understory trees of a
Central Amazon forest during the four field seasons of the 2015 El Niño year. Boxplot and
ANOVA details are as described in Fig. 3.
Fig. 6 Box plots of leaf nitrogen (N; a), phosphorus (P; b) and potassium (K; c) concentration
in upper canopy and understory trees of a Central Amazon forest during the four field seasons
of the 2015 El Niño year. Boxplot and ANOVA details are as described in Fig. 3.
Fig. 7 Correlations between light saturated photosynthesis (Asat) and stomatal conductance
(gs), using their respective relative differences () between the extreme dry season of 2015
( Dry) and the mean of the other three seasons ( Annual), where  Dry/ Annual)-1. (ncanopy =
15; nunderstory = 24).
Accepted Article
This article is protected by copyright. All rights reserved.
Supporting Information captions
Table S1. Canopy and understory species list
Table S2. Univariate test for repeated measure (season)

Figure S4. Box plot of maximum carboxylation rate of ribulose-1, 5-bisphosphate
carboxylase/oxygenase (Vcmax) seasonality
Supplementary Text S5. Accumulated Daily Water Deficit (ADWD) and Relative Extractable
Water (REW)
Figure S6. Two years of Accumulated Daily Water Deficit, Daily Rainfall, and Relative
Extractable Water
Figure S7. Nineteen years of Accumulated Daily Water Deficits
Figure S8. A_PPFD response curve
Figure S9. Box plot of predawn and mid-day leaf water potential
Supplementary Text S10. Repeated-measures ANOVA to detect effects of season and of
plant family
Table S11. Repeated Measures Analysis of Variance for understory trees
Figure S12. Photosynthesis seasonality in the understory for four tree families
Table S13. Repeated Measures Analysis of Variance for upper canopy trees
Figure S14. Photosynthesis seasonality in the upper canopy for four tree families.
Supplementary Text S15 Repeated-measures ANOVA to detect effects of season and of light
environment
Table S16. Repeated Measures Analysis of Variance
Figure S17. Photosynthesis seasonality for the four light levels from Dawkins index.
Excel worksheet S18. Data
Accepted Article
This article is protected by copyright. All rights reserved.
Accepted Article
This article is protected by copyright. All rights reserved.
Accepted Article
This article is protected by copyright. All rights reserved.
Accepted Article
This article is protected by copyright. All rights reserved.
Accepted Article
This article is protected by copyright. All rights reserved.
Accepted Article
This article is protected by copyright. All rights reserved.
Accepted Article
This article is protected by copyright. All rights reserved.
... Do not consider these interactions to influence substantial errors in climate measurements, due to temperature variation between the ground surface and the air masses above the canopy [59]. Thus, this study provides direct evidence of climate change, as it presents the effects of El Niño in the Amazon region, as well as its repercussions for the region's ecosystem [60][61][62], especially the low availability of pasture as a source of food for animals. ...
Article
Full-text available
Simple Summary The El Niño presents itself as a serious problem in the pastures of the northern region of Brazil, as it compromises the availability and quality of forage and water. Therefore, the objective of this study was to characterize the thermographic profile of three production systems in the Eastern Amazon, Brazil. The results show significant differences between areas with and without chestnut tree shade. Between August and November, the highest temperatures were recorded in full sun pastures, contrasting with lower temperatures in shaded areas. The interaction between the systems revealed significant thermal variations, highlighting the positive impact of native trees on thermal regulation and indicating possible strategies to mitigate the adverse effects of El Niño. Abstract The El Niño represents a substantial threat to pastures, affecting the availability of water, forage and compromising the sustainability of grazing areas, especially in the northern region of Brazil. Therefore, the objective of this study was to characterize the thermographic profile of three production systems in the Eastern Amazon, Brazil. The study was conducted on a rural cattle farm in Mojuí dos Campos, Pará, Brazil, between August and November 2023. The experiment involved livestock production systems, including traditional, silvopastoral and integrated, with different conditions of shade and access to the bathing area. An infrared thermographic (IRT) camera was used, recording temperatures in different zones, such as areas with trees, pastures with forage and exposed pastures. The highest mean temperatures (p = 0.02) were observed in pastures with full sun from August to November. On the other hand, the lowest average temperatures were recorded in areas shaded by chestnut trees (Bertholletia excelsa). The highest temperature ranges were found in sunny areas and the lowest were recorded in shaded areas. The highest temperatures were observed in the pasture in full sun, while the lowest were recorded in areas shaded by chestnut trees (p < 0.0001). The interaction between the systems and treatments revealed significant temperature differences (p < 0.0001), with the native trees showing an average temperature of 35.9 °C, lower than the grasses and soil, which reached 61.2 °C. This research concludes that, under El Niño in the Eastern Amazon, areas shaded by Brazil nut trees had lower temperatures, demonstrating the effectiveness of shade. Native trees, compared to grasses and soil, showed the ability to create cooler environments, highlighting the positive influence on different species such as sheep, goats and cattle.
... Maximum monthly rates of precipitation occur in March/April (> 300 mm), and the period between June and October is usually drier but still with an average of more than 100 mm of rainfall per month. However, rainfall can be below 100 mm between June and October in exceptionally dry years (Kunert et al., 2015;Santos et al., 2018). The average air temperature is 25.8 • C with little variation between seasons. ...
Article
Tropical forests are important carbon sinks as they store huge amounts of carbon and are thus essential players within the global carbon cycle. While good estimates of the gross and net primary productivity (GPP and NPP, respectively) of tropical forests exist, we still lack a cohesive study on the contribution of different tree sizes to the overall forest carbon fluxes and their seasonal variation. Here, we model GPP based on xylem sap flux and eddy covariance data for an old-growth moist lowland forest in Central Amazon for the year 2013 (Jan-Dec). The model uses canopy transpiration and vapor pressure deficit as covariates and enables GPP partitioning into fluxes from canopy, subcanopy, and understory trees. Net primary productivity was calculated from forest inventory data. Carbon use efficiency (CUE) was computed as the ratio between NPP and GPP. GPP was 28.46 MgC ha − 2 yr − 1 at our study site. Canopy trees (diameter > 30 cm; average height of 28 m) were responsible for 21.47 MgC ha − 2 yr − 1 of the overall GPP, whereas subcanopy and understory trees contributed 3.95 MgC ha − 2 yr − 1 and 3.04 MgC ha − 2 yr − 1 , respectively. Canopy trees allocated only 23% of their photosynthetic products towards growth and were responsible for more than 50% of NPP. Subcanopy and understory trees were more efficient by allocating 67% and 59% of their carbon assimilates towards biomass growth, respectively. GPP showed distinct seasonal patterns, with canopy trees doubling monthly GPP with the progressing dry season, whereas understory and subcanopy trees had a 60% higher GPP than in the wet season. This study provides evidence for the importance of large trees in tropical forests and highlights their crucial role in forest carbon cycling. Due to high drought-related mortality, large trees make up a critical component of the response of tropical forests to climate change.
... Petchiappan et al. (2022) (Goncalves et al., 2020). As for the possible causes for the quick 290 recovery of photosynthetic rate, Santos et al. (2018) found that the photosynthesis reduction under extreme drought and high temperature in the 2015 dry season was primarily due to stomatal closure, which can reverse when water becomes available. ...
Preprint
Full-text available
The 2015/16 Amazon drought was characterized by below-average regional precipitation for an entire year, which distinguishes it from the dry-season only droughts in 2005 and 2010. Studies of vegetation indices (VI) derived from optical remote sensing over the Amazonian forests indicated three stages in canopy response during the 2015/16 drought, with below-average greenness during the onset and end of the drought, and above-average greenness during the intervening months. So far, a satisfactory explanation for this broad temporal pattern, and spatial variation within the Amazonian forests of this broad response, has not been found. Better understanding of rainforest behaviors during this unusually long drought should help predict their response to future droughts. We hypothesized that below-average greenness could be explained by water deficit and heat stress occurring beyond the tolerance thresholds of rainforest. To test our hypothesis, we used monthly observations of terrestrial water storage (TWS), land surface temperature (LST) and vapor pressure deficit (VPD) for January 2003–December 2016. First, for each 1° grid cell, we determined the ‘normal’ range of monthly TWS, LST and VPD during non-drought years (i.e. 2003–2016, excluding 2005, 2010, 2015 and 2016), and identified the extreme values of ‘normal’ range, i.e. minimum TWS, maximum LST and maximum VPD. We considered the normal hydrological and thermal ranges to have been exceeded when (1) two or three of these variables were simultaneously beyond their extreme values, or (2) only one variable was beyond the extreme value, but the other two were significantly (p<0.05) different from the average for non-drought years. Using these criteria, regions experiencing hydrological and thermal conditions beyond the ‘normal’ range during different stages of the 2015/16 event were delineated. The results showed a gradual southward shift of these regions: from the north-eastern Amazon in August–October 2015, to the north-central part in November 2015–February 2016 and finally to the southern Amazon in July 2016. The majority of forests within the delimited regions experienced below-average greenness. Conversely, outside of these regions, greenness responded positively to radiation anomalies, as is expected under normal conditions. The opposing influences of drought and radiation anomalies together explained more than 70 % of the observed spatiotemporal patterns in greenness. These results suggest that our exceeding ‘normal’ ranges based approach, combining water storage, temperature and atmospheric moisture demand drivers, can reasonably identify the most likely drought-affected regions at monthly to seasonal time scales. Using observation-based hydrological and thermal condition thresholds can help with interpreting the response of Amazon rainforest to future drought events.
... injured the photosynthetic apparatus and diminished the efficiency of light energy utilization (Meng et al., 2016;Santos et al., 2018;Tantray et al., 2020). The limitation of carbon assimilation capacity is the primary factor contributing to the decline in Naidong growth following drought conditions. ...
Article
Due to population growth and expanding economic activities, waste generation is on the increase globally. Reuse of waste or by-products in agriculture as biostimulants and fertilizers contributes to a circular economy. In recent years, there has been an increasing number of studies on organic waste-derived biostimulants and the economic market for these products is expanding. Many studies have addressed the development of biostimulants from various organic wastes or by-products. It is worth noting that different types of organic wastes or by-products exhibit different compositions and key mechanisms when used as biostimulants, which require analyzing the mechanism of action from multiple fields such as chemistry and biology. Therefore, this review presents recent advances with respect to organic waste-derived biostimulants. Specifically, it outlines the characteristics of biostimulants derived from different waste or by-products and extraction techniques. The role of organic-waste derived biostimulants in maintaining plant growth and development is described, emphasizing their role in helping plants resist several abiotic stresses. Mechanisms by which organic waste-derived biostimulants affect physiological changes in plants are discussed. This review provides the theoretical foundation to enable the efficient waste material and by-products resource utilization and evidences that the development and production of biostimulants from waste and by-products is an environmentally friendly option for waste recycling and for the increase of crop productivity.
Preprint
Full-text available
A large fraction of the interannual variations in the global carbon cycle can be explained and predicted by the impact of El Niño Southern Oscillation (ENSO) on net biome production (NBP). It is therefore crucial that the relationship between ENSO and NBP is correctly represented in Earth system model (ESMs). With this work, we look beyond the top-down ENSO-CO2 relationship in 22 CMIP6 ESMs by describing their characteristic ENSO-NBP pathways. These pathways result from the configuration of three interacting processes which contribute to the overall ENSO-CO2 relationship: ENSO-strength, ENSO-induced climate anomalies, and the sensitivity of NBP to climate. The analysed ESMs agree on the direction of the sensitivity of global NBP to ENSO, but have very large uncertainty in its magnitude, with a global NBP anomaly of -0.15 PgC yr-1 to -2.13 PgC yr-1 per standardised El Niño event. The largest source of uncertainty is the differences in the sensitivity of NBP to climate. The uncertainty among the ESMs grows even further when only the differences in NBP sensitivity to climate are considered. This is because differences in the climate sensitivity of NBP are partially compensated by ENSO strength. There is a similar phenomenon regarding the distribution of ENSO-induced climate anomalies. We show that even model that agree on global NBP anomalies have strong disagreements in the contribution of different regions to the global anomaly. This analysis shows, that while ESMs can have a comparable ENSO-induced CO2 anomaly, the carbon fluxes contributing to this anomaly originate from different regions and are caused by different drivers. The consequence of these alternative ENSO-NBP pathways can be a false confidence in the reproduction of CO2 by assimilating the ocean, and the dismissal of predictive performance offered through ENSO. We suggest to improve the underlying processes by using large-scale carbon flux data for model tuning in order to capture the ENSO-induced NBP anomaly patterns. The increasing availability of carbon flux data from atmospheric inversions and remote sensing products makes this a tangible goal and would lead to a better representation of the processes driving the interannual variability of the global carbon cycle.
Article
Full-text available
Soil microbiota can confer fitness advantages to plants and increase crop resilience to drought and other abiotic stressors. However, there is little evidence on the mechanisms correlating a microbial trait with plant abiotic stress tolerance. Here, we report that Streptomyces effectively alleviate drought and salinity stress by producing spiroketal polyketide pteridic acid H (1) and its isomer F (2), both of which promote root growth in Arabidopsis at a concentration of 1.3 nM under abiotic stress. Transcriptomics profiles show increased expression of multiple stress responsive genes in Arabidopsis seedlings after pteridic acids treatment. We confirm in vivo a bifunctional biosynthetic gene cluster for pteridic acids and antimicrobial elaiophylin production. We propose it is mainly disseminated by vertical transmission and is geographically distributed in various environments. This discovery reveals a perspective for understanding plant-Streptomyces interactions and provides a promising approach for utilising beneficial Streptomyces and their secondary metabolites in agriculture to mitigate the detrimental effects of climate change.
Article
Full-text available
How tropical plants cope with water availability has important implications for forest resilience, as severe drought events are expected to increase with climate change. Tree size has emerged as a major axis of drought vulnerability. To understand how Amazon tree species are distributed along size-linked gradients of water and light availability, we tested the niche acclimation hypothesis that there is a developmental gradient in ontogenetic shift in embolism resistance and tree water-use efficiency among tree species that occurs along the understory-overstory gradient. We evaluated ontogenetic differences in the intrinsic water-use efficiency (iWUE) and xylem hydraulic traits of abundant species in a seasonal tropical forest in Brazil. We found that saplings of dominant overstory species start with a high degree of embolism resistance to survive in a dense understory environment where competition for water and light among smaller trees can be intense during the prolonged dry season. Vulnerability to embolism consistently changed with ontogeny and varied with tree species' stature (maximum height): mature individuals of larger species displayed increased vulnerability, whereas smaller species displayed unchanging or even increased resistance at the mature stage. The ability to change drought-resistance strategies (vulnerability to embolism) through ontogeny was positively correlated with ontogenetic increase in iWUE. Ecologically, overstory trees appear to shift from being hydraulically drought resilient to persisting under dry soil surface layer conditions to being more likely physiological drought avoiders as adults when their roots reach wetter and deeper soil layers.
Article
Interspecific variations in phenotypic plasticity of trees that are affected by climate change may alter the ecosystem function of forests. Seedlings of four common tree species (Castanopsis fissa, Michelia macclurei, Dalbergia odorifera and Ormosia pinnata) in subtropical plantations of southern China were grown in the field under rainout shelters and subjected to changing precipitation (48 L of water every 4 days in the dry season, 83 L of water every 1 day in the wet season; 4 g m−2 year−1 of nitrogen (N)), low N deposition (48 L of water every 2 days in the dry season, 71 L of water every 1 day in the wet season; 8 g m−2 year−1 N), high N deposition (48 L of water every 2 days in the dry season, 71 L of water every 1 day in the wet season; 10 g m−2 year−1 N) and their interactive effects. We found that the changes in seasonal precipitation reduced the light-saturated photosynthetic rate (Asat) for C. fissa due to declining area-based foliar N concentrations (Na). However, we also found that the interactive effects of changing precipitation and N deposition enhanced Asat for C. fissa by increasing foliar Na concentrations, suggesting that N deposition could alleviate N limitations associated with changing precipitation. Altered precipitation and high N deposition reduced Asat for D. odorifera by decreasing the maximum electron transport rate for RuBP regeneration (Jmax) and maximum rate of carboxylation of Rubisco (Vcmax). Ormosia pinnata under high N deposition exhibited increasing Asat due to higher stomatal conductance and Vcmax. The growth of D. odorifera might be inhibited by changes in seasonal precipitation and N deposition, while O. pinnata may benefit from increasing N deposition in future climates. Our study provides an important insight into the selection of tree species with high capacity to tolerate changing precipitation and N deposition in subtropical plantations.
Article
Carbon dioxide and water vapor exchanges between tropical forest canopies and the atmosphere through photosynthesis, respiration, and evapotranspiration (ET) influence carbon and water cycling at the regional and global scales. Their inter- and intra-annual variations are sensitive to seasonal rhythms and longer-timescale tropical climatic events. In the present study, we assessed the El Niño-Southern Oscillation (ENSO) influence on ET and on the net ecosystem exchange (NEE), using eddy-covariance flux observations in a Bornean rainforest over a 10-y period (2010–2019) that included several El Niño and La Niña events. From flux model inversions, we inferred ecophysiological properties, notably the canopy stomatal conductance and “big-leaf” maximum carboxylation rate ( V cmax25_BL ). Mean ET values were similar between ENSO phases (El Niño, La Niña, and neutral conditions). Conversely, the mean net ecosystem productivity was highest during La Niña events and lowest during El Niño events. Combining Shapley additive explanation calculations for nine controlling factors with a machine-learning algorithm, we determined that the primary factors for ET and NEE in the La Niña and neutral phases were incoming shortwave solar radiation and V cmax25_BL , respectively, but that canopy stomatal conductance was the most significant factor for both ET and NEE in the El Niño phase. A combined stomatal-photosynthesis model approach further indicated that V cmax25_BL differences between ENSO phases were the most significant controlling factor for canopy photosynthesis, emphasizing the strong need to account for ENSO-induced ecophysiological factor variations in model projections of the long-term carbon balance in Southeast Asian tropical rainforests.
Article
Full-text available
Net photosynthetic carbon uptake of Panamanian lowland tropical forest species is typically optimal at 30–32°C. The processes responsible for the decrease in photosynthesis at higher temperatures are not fully understood for tropical trees. We determined temperature responses of maximum rates of RuBP-carboxylation (VCMax) and RuBP-regeneration (JMax), stomatal conductance (Gs) and respiration in the light (RLight) in situ for four lowland tropical tree species in Panama. Gs had the lowest temperature optimum (TOpt), similar to that of net photosynthesis, and photosynthesis became increasingly limited by stomatal conductance as temperature increased. JMax peaked at 34–37°C and VCMax ~2°C above that, except in the late-successional species Calophyllum longifolium, in which both peaked at ~33°C. RLight significantly increased with increasing temperature, but simulations with a photosynthesis model indicated that this had only a small effect on net photosynthesis. We found no evidence for Rubisco-activase limitation of photosynthesis. TOpt of VCMax and JMax fell within the observed in situ leaf temperature range, but our study nonetheless suggests that net photosynthesis of tropical trees is more strongly influenced by the indirect effects of high temperature—e.g., through elevated vapor pressure deficit and resulting decreases in stomatal conductance—than by direct temperature effects on photosynthetic biochemistry and respiration.
Article
Full-text available
Changes in tropical forest carbon sink strength during El Niño Southern Oscillation ( ENSO ) events can indicate future behavior under climate change. Previous studies revealed ˜6 Mg C ha ⁻¹ yr ⁻¹ lower net ecosystem production ( NEP ) during ENSO year 1998 compared with non‐ ENSO year 2000 in a Costa Rican tropical rainforest. We explored environmental drivers of this change and examined the contributions of ecosystem respiration ( RE ) and gross primary production ( GPP ) to this weakened carbon sink. For 1998–2000, we estimated RE using chamber‐based respiration measurements, and we estimated GPP in two ways: using (1) the canopy process model MAESTRA , and (2) combined eddy covariance and chamber respiration data. MAESTRA ‐estimated GPP did not statistically differ from GPP estimated using approach 2, but was ˜ 28% greater than published GPP estimates for the same site and years using eddy covariance data only. A 7% increase in RE (primarily increased soil respiration) and a 10% reduction in GPP contributed equally to the difference in NEP between ENSO year 1998 and non‐ ENSO year 2000. A warming and drying climate for tropical forests may yield a weakened carbon sink from both decreased GPP and increased RE . Understanding physiological acclimation will be critical for the large carbon stores in these ecosystems.
Article
Full-text available
Our limited understanding of the climate controls on tropical forest seasonality is one of the biggest sources of uncertainty in modeling climate change impacts on terrestrial ecosys- tems. Combining leaf production, litterfall and climate observations from satellite and ground data in the Amazon forest, we show that seasonal variation in leaf production is largely trig- gered by climate signals, specifically, insolation increase (70.4% of the total area) and pre- cipitation increase (29.6%). Increase of insolation drives leaf growth in the absence of water limitation. For these non-water-limited forests, the simultaneous leaf flush occurs in a suffi- cient proportion of the trees to be observed from space. While tropical cycles are generally defined in terms of dry or wet season, we show that for a large part of Amazonia the increase in insolation triggers the visible progress of leaf growth, just like during spring in temperate forests. The dependence of leaf growth initiation on climate seasonality may result in a higher sensitivity of these ecosystems to changes in climate than previously thought.
Article
Full-text available
Climate change exposes vegetation to unusual levels of drought, risking a decline in productivity and an increase in mortality. It still remains unclear how trees and forests respond to such unusual drought, particularly Southeast Asian tropical rain forests. To understand leaf ecophysiological responses of tropical rain forest trees to soil drying, a rainfall exclusion experiment was conducted on mature canopy trees of Dryobalanops aromatica Gaertn.f. (Dipterocarpaceae) for 4 months in an aseasonal tropical rain forest in Sarawak, Malaysia. The rainfall was intercepted by using a soft vinyl chloride sheet. We compared the three control and three treatment trees with respect to leaf water use at the top of the crown, including stomatal conductance (gsmax), photosynthesis (Amax), leaf water potential (predawn: Ψpre; midday: Ψmid), leaf water potential at turgor loss point (πtlp), osmotic potential at full turgor (π100) and a bulk modulus of elasticity (ε). Measurements were taken using tree-tower and canopy-crane systems. During the experiment, the treatment trees suffered drought stress without evidence of canopy dieback in comparison with the control trees; e.g., Ψpre and Ψmid decreased with soil drying. Minimum values of Ψmid in the treatment trees decreased during the experiment, and were lower than πtlp in the control trees. However, the treatment trees also decreased their πtlp by osmotic adjustment, and the values were lower than the minimum values of their Ψmid. In addition, the treatment trees maintained gs and Amax especially in the morning, though at midday, values decreased to half those of the control trees. Decreasing leaf water potential by osmotic adjustment to maintain gs and Amax under soil drying in treatment trees was considered to represent anisohydric behavior. These results suggest that D. aromatica may have high leaf adaptability to drought by regulating leaf water consumption and maintaining turgor pressure to improve its leaf water relations.
Article
Full-text available
• The photosynthetic performance of tropical forests in a warming Earth is uncertain. To decrease this uncertainty, it is critical and necessary to gain a better understanding of the optimum temperature for photosynthesis (Topt) and the photosynthetic response to warming. • With the aid of ecosystem flux data for seven tropical forests across different continents, we quantified ecosystem Topt (ToptE). Photosynthetic parameters were inverted after abstracting each forest into a big-leaf. These parameters were used to separate the contributions of biochemical, respiratory, and stomatal processes in determining ToptE. • The ecosystem ToptE was close to the mean growth temperature and tended to increase with growth temperature across sites. Respiratory and stomatal processes but not biochemical processes played important roles in ToptE changes. ToptE increased from approximately 26 °C to above 30 °C when the water vapour deficit was controlled. • Stomatal processes showed a disproportional importance in determining ToptE. The apparent ecosystem ToptE may plausibly increase with warming through thermal acclimation of respiration and stomatal sensitivity. Similarly, our results indicate that increases in ambient CO2 increase ToptE by alleviating stomatal limitation on photosynthesis under high water pressure deficit.
Article
Full-text available
Accurate estimation of terrestrial gross primary productivity (GPP) remains a challenge despite its importance in the global carbon cycle. Chlorophyll fluorescence (ChlF) has been recently adopted to understand photosynthesis and its response to the environment, particularly with remote sensing data. However, it remains unclear how ChlF and photosynthesis are linked at different spatial scales across the growing season. We examined seasonal relationships between ChlF and photosynthesis at the leaf, canopy, and ecosystem scales, and explored how leaf-level ChlF was linked with canopy-scale solar induced chlorophyll fluorescence (SIF) in a temperate deciduous forest at Harvard Forest, Massachusetts, USA. Our results show that ChlF captured the seasonal variations of photosynthesis with significant linear relationships between ChlF and photosynthesis across the growing season over different spatial scales (R(2) =0.73, 0.77 and 0.86 at leaf, canopy and satellite scales, respectively; p<0.0001). We developed a model to estimate GPP from the tower-based measurement of SIF and leaf-level ChlF parameters. The estimation of GPP from this model agreed well with flux tower observations of GPP (R(2) =0.68; p<0.0001), demonstrating the potential of SIF for modeling GPP. At the leaf scale, we found that leaf Fq '/Fm ', the fraction of absorbed photons that are used for photochemistry for a light adapted measurement from a pulse amplitude modulation fluorometer, was the best leaf fluorescence parameter to correlate with canopy-SIF yield (SIF/APAR, R(2) =0.79; p<0.0001). We also found that canopy-SIF and SIF-derived GPP (GPPSIF ) were strongly correlated to leaf-level biochemistry and canopy structure, including chlorophyll content (R(2) =0.65 for canopy-GPPSIF and chlorophyll content; p<0.0001), leaf area index (LAI) (R(2) =0.35 for canopy-GPPSIF and LAI; p<0.0001), and normalized difference vegetation index (NDVI) (R(2) =0.36 for canopy-GPPSIF and NDVI; p<0.0001). Our results suggest that ChlF can be a powerful tool to track photosynthetic rates at leaf, canopy, and ecosystem scales. This article is protected by copyright. All rights reserved.
Article
Full-text available
The El Niño-Southern Oscillation (ENSO) is the main driver of interannual climate extremes in Amazonia and other tropical regions. The current 2015/2016 EN event was expected to be as strong as the EN of the century in 1997/98, with extreme heat and drought over most of Amazonian rainforests. Here we show that this protracted EN event, combined with the regional warming trend, was associated with unprecedented warming and a larger extent of extreme drought in Amazonia compared to the earlier strong EN events in 1982/83 and 1997/98. Typical EN-like drought conditions were observed only in eastern Amazonia, whilst in western Amazonia there was an unusual wetting. We attribute this wet-dry dipole to the location of the maximum sea surface warming on the Central equatorial Pacific. The impacts of this climate extreme on the rainforest ecosystems remain to be documented and are likely to be different to previous strong EN events.
Article
Full-text available
The seasonal climate drivers of the carbon cycle in tropical forests remain poorly known, although these forests account for more carbon assimilation and storage than any other terrestrial ecosystem. Based on a unique combination of seasonal pan-tropical data sets from 89 experimental sites (68 include aboveground wood productivity measurements and 35 litter productivity measurements), their associated canopy photosynthetic capacity (enhanced vegetation index, EVI) and climate, we ask how carbon assimilation and aboveground allocation are related to climate seasonality in tropical forests and how they interact in the seasonal carbon cycle. We found that canopy photosynthetic capacity seasonality responds positively to precipitation when rainfall is < 2000 mm yr(-1) (water-limited forests) and to radiation otherwise (light-limited forests). On the other hand, independent of climate limitations, wood productivity and litterfall are driven by seasonal variation in precipitation and evapotranspiration, respectively. Consequently, light-limited forests present an asynchronism between canopy photosynthetic capacity and wood productivity. First-order control by precipitation likely indicates a decrease in tropical forest productivity in a drier climate in water-limited forest, and in current light-limited forest with future rainfall < 2000 mm yr(-1).
Article
Chlorophyll (Chl) a fluorescence induction (transient), measured by exposing dark-adapted samples to high light, shows a polyphasic rise, which has been the subject of extensive research over several decades. Several Chl fluorescence parameters based on this transient have been defined, the most widely used being the FV [= (FM–F0)]/FM ratio as a proxy for the maximum quantum yield of PSII photochemistry. However, considerable additional information may be derived from analysis of the shape of the fluorescence transient. In fact, several performance indices (PIs) have been defined, which are suggested to provide information on the structure and function of PSII, as well as on the efficiencies of specific electron transport reactions in the thylakoid membrane. Further, these PIs have been proposed to quantify plant tolerance to stress, such as by high light, drought, high (or low) temperature, or N-deficiency. This is an interesting idea, since the speed of the Chl a fluorescence transient measurement (<1 s) is very suitable for high-throughput phenotyping. In this review, we describe how PIs have been used in the assessment of photosynthetic tolerance to various abiotic stress factors. We synthesize these findings and draw conclusions on the suitability of several PIs in assessing stress responses. Finally, we highlight an alternative method to extract information from fluorescence transients, the Integrated BiomarkerResponse. This method has been developed to define multi-parametric indices in other scientific fields (e.g., ecology), and may be used to combine Chl a fluorescence data with other proxies characterizing CO2 assimilation, or even growth or grain yield, allowing a more holistic assessment of plant performance.