Skip to main content
Journal of Experimental Botany logoLink to Journal of Experimental Botany
. 2023 Nov 4;75(3):962–978. doi: 10.1093/jxb/erad437

Photosynthesis in newly developed leaves of heat-tolerant wheat acclimates to long-term nocturnal warming

Onoriode Coast 1,2,3,, Andrew P Scafaro 4,5, Helen Bramley 6,b, Nicolas L Taylor 7, Owen K Atkin 8,9
Editor: Elizabete Carmo-Silva10
PMCID: PMC10837020  PMID: 37935881

Abstract

We examined photosynthetic traits of pre-existing and newly developed flag leaves of four wheat genotypes grown in controlled-environment experiments. In newly developed leaves, acclimation of the maximum rate of net CO2 assimilation (An) to warm nights (i.e. increased An) was associated with increased capacity of Rubisco carboxylation and photosynthetic electron transport, with Rubisco activation state probably contributing to increased Rubisco activity. Metabolite profiling linked acclimation of An to greater accumulation of monosaccharides and saturated fatty acids in leaves; these changes suggest roles for osmotic adjustment of leaf turgor pressure and maintenance of cell membrane integrity. By contrast, where An decreased under warm nights, the decline was related to lower stomatal conductance and rates of photosynthetic electron transport. Decreases in An occurred despite higher basal PSII thermal stability in all genotypes exposed to warm nights: Tcrit of 45–46.5 °C in non-acclimated versus 43.8–45 °C in acclimated leaves. Pre-existing leaves showed no change in An–temperature response curves, except for an elite heat-tolerant genotype. These findings illustrate the impact of night-time warming on the ability of wheat plants to photosynthesize during the day, thereby contributing to explain the impact of global warming on crop productivity.

Keywords: Acclimation, electron transport thermotolerance, high night temperature, photosynthetic efficiency, Triticum aestivum


Wheat photosynthesis and acclimation capacity are directly affected by night warming, despite the temporal separation of night warming from photosynthesis, with varying responses between newly developed and pre-existing leaves.

Introduction

Wheat (Triticum aestivum L) is the most widely cultivated crop, with >220 Mha cultivated worldwide. Wheat yield must increase by 0.83% per annum to meet the projected 44% increase in demand for it by 2050 (CGIAR Advisory Services, 2020). To achieve an environmentally sustainable increase in production of this magnitude, while also coping with rising global surface temperatures, represents a major challenge. Moreover, most research has focused on effects of elevated temperatures during the day on wheat, but rising temperatures have been more pronounced for daily minimums (i.e. night warming) than for daily maximums (Easterling et al., 1997; Alexander et al., 2006; Sillmann et al., 2013). Night warming is linked with a reduction in wheat yield in many regions including the USA, Mexico, and Argentina (Lobell et al., 2005; Prasad et al., 2008; García et al., 2015, 2016, 2018; Russell and Van Sanford, 2020; Fisher et al., 2022). Given the importance of wheat for global food production, it is vital that we develop a more complete understanding of the physiological basis for why nocturnal warming reduces wheat yields. Reflecting this, interest in wheat physiological responses to warm nights has increased, with reports on steady-state photosynthesis, acclimation of leaf and root respiration, and altered source–sink capacity at anthesis and grain filling (Fan et al., 2015, 2017; Impa et al., 2019, 2020; Bahuguna et al., 2022; Posch et al., 2022a). However, there remain key gaps in knowledge, especially in relation to the effect of warm nights on overall photosynthetic capacity, and whether key photosynthetic processes such as the critical temperature above which incipient damage to photosystem II (PSII) occurs (Tcrit) acclimate to warm nights. PSII is considered the most heat-sensitive component of electron transport in the photosynthetic machinery, with heat stress resulting in the unfolding of protein complexes and loss of manganese from the oxygen-evolving complex (Enami et al., 1994).

According to the Farquhar, von Caemmerer, and Berry (FvCB model; Farquhar et al., 1980; von Caemmerer, 2000) biochemical model of photosynthesis, light-saturated An at a given temperature can be limited by: (i) the maximum capacity of Rubisco carboxylation (Vcmax) and internal CO2 concentration (Ci); or (ii) the regeneration of ribulose 1,5-bisphosphate (RuBP) regeneration which relies on photosynthetic electron transport capacity (J) and irradiance. An can also be limited if triose phosphate use (TPU) restricts the recycling of phosphate within the chloroplast (Sharkey, 1985). Vcmax, J, and TPU can be estimated by fitting the FvCB model (Farqhuar et al., 1980) to plots of light-saturated An versus Ci. Light-saturated rates of An respond to short- and long-term changes in growth environment temperature (Atkin and Tjoelker, 2003; Way and Yamori, 2014). The thermal response of light-saturated An to short-term changes in growth environment temperature is curvilinear—increasing with a rise in temperature at suboptimal temperatures to a maximum rate (Amax) at the optimum temperature (Topt), then decreasing at temperatures beyond Topt. Longer term changes in the thermal environment cause either positive or detractive adjustments of light-saturated An. Positive adjustments increase Amax and/or shift Topt towards the new, higher growth temperature, indicating acclimation, whereas detractive adjustments reduce Amax and/or Topt (Sage and Kubien, 2007; Way and Yamori, 2014; Posch et al., 2019). Changes in the An–temperature response are controlled by different factors, including the temperature sensitivities of Vcmax and J (Hikosaka et al., 2005; Sage and Kubien, 2007; Lin et al., 2012), with alterations in Vcmax reflecting changes in the abundance and/or activation state of Rubisco (Scafaro et al., 2023). Another controlling factor, albeit indirectly, is stomatal conductance (gs). At high temperatures, stomatal closure reduces water loss when the atmospheric vapour pressure deficit (VPD) rises, but this simultaneously reduces CO2 uptake. In wheat, the An–temperature response to elevated daily temperatures varies from positive acclimation to detractive adjustments (Yamasaki et al., 2002; Cai et al., 2018, 2020; Coast et al., 2021a). Importantly, information on thermal acclimation of wheat photosynthesis to specifically warm nights is scarce (but see Turnbull et al., 2002 for work on a broadleaved tree species). It is likely that under night warming, wheat photosynthesis and acclimation capacity may be affected directly by constraining leaf biochemical components underlying photosynthesis (i.e. Vcmax and J), or indirectly by stomatal closure, or both. These suggestions remain untested.

The degree of physiological acclimation in leaves to warming varies substantially with the duration and intensity of warming, as well as the plant’s developmental stage. Leaves that develop under a new warming regime exhibit a higher degree of acclimation than pre-existing leaves that were formed under an earlier thermal regime and which then experience a sustained change in growth temperature (Loveys et al., 2003; Yamori et al., 2005; Rashid et al., 2020; Slot et al., 2021). Varying acclimation capacity with leaf development has also been shown in cold regimes (Hurry et al., 2005; Campbell et al., 2007). Limited reports exist about leaf development under warm conditions and the magnitude of photosynthetic acclimation in newly developed and pre-existing leaves. Determining these responses could be relevant in managing crop performance and useful in parameterizing carbon fluxes in Earth system models (Rogers et al. 2017).

T crit, a common measure of photosynthetic heat tolerance (Geange et al., 2020; Lancaster and Humphreys, 2020), is estimated from temperature-dependent changes in minimum Chl a fluorescence (F0) (Schreiber et al., 1975; Schreiber and Berry, 1977). Tcrit correlates with the maximum temperature at which An approaches zero (Downton et al., 1984). Under warmer growth conditions, an increase in Tcrit is considered evidence of acclimation and indicative of higher temperature tolerance. Acclimation of Tcrit to warmer growth conditions has been shown in many non-crop, mostly woody species. For example, 62 Australian species acclimated seasonally with Tcrit, increasing on average from winter to summer by 0.34 °C per °C increase in growth temperature (Zhu et al., 2018). Similar reports exist for Mediterranean and temperate European oak species (Quercus spp., Daas et al., 2008), and at a global scale for 218 plant species spanning seven biomes (O’Sullivan et al., 2017). Acclimation of Tcrit is assumed to increase the upper thermal threshold of photosynthesis by relaxing the limits of photosynthetic electron transport. As such, the capacity to acclimate Tcrit might underlie photosynthetic performance under warm nights. However, it is not known whether Tcrit acclimates in response to warm nights.

Metabolites function as regulatory agents, compatible solutes, antioxidants, and reductants in adaptation to thermal stress. Warming alters the abundance of primary metabolites [sugars, sugar alcohols, organic acids, amino acids, or tricarboxylic acid (TCA) cycle intermediates] in leaves of stress-sensitive cereals (Glaubitz et al., 2014; Impa et al., 2019; Schaarschmidt et al., 2020; Rashid et al., 2021). Increases in the abundance of organic acids and TCA cycle components in heat-sensitive genotypes under heat stress often reflect impairment to CO2 assimilation, regulation of the TCA cycle, and amino acid biosynthesis, and these processes are central to plant photosynthesis. Deciphering changes in the metabolic phenotype of wheat in response to warm nights could aid our understanding of plant responses to climate change and provide input for developing adaptation tools for crop production in a warmer world.

To address some of the above issues, we compared four wheat genotypes (including a commercial Australian cultivar and an elite heat-tolerant genotype) under night temperatures of 15, 20, or 25 °C for different durations (5–7 d at anthesis or 9–13 weeks prior to and including anthesis). Our primary objectives were to: (i) quantify the acclimation capacity of photosynthesis and Tcrit to warm nights; and (ii) assess changes in the wheat metabolic phenotype under warm nights. In addition, we set out to compare effects of warm nights on the temperature dependencies of photosynthetic parameters in newly developed versus pre-existing leaves. The results indicate that plants stressed by warm nights reduced photosynthetic performance via down-regulation of J even while exhibiting high Tcrit (i.e. high PSII thermal stability). By contrast, tolerance of warm nights (in terms of photosynthesis) was marked by improved An linked to acclimation in Vcmax, PSII thermal stability, and an increase in leaf metabolic signatures for monosaccharides.

Materials and methods

To assess the extent to which photosynthetic capacity of pre-existing and newly developed leaves of wheat respond to warm nights, two experiments were conducted. In Experiment I, acclimation of Tcrit, instantaneous temperature response functions of photosynthetic capacity [light-saturated An, Vcmax, J1500 (J measured at a photosynthetic photon flux of 1500 µmol m–2 s–1), and TPU], and changes in the metabolite profile of plants that developed from seedling to anthesis (9–13 weeks of night warming) were compared at three night temperatures (15, 20, and 25 °C) and a common day temperature of 26 °C. These temperature treatments are reflective of conditions in some wheat-growing regions including parts of the USA (Hein et al., 2020) and India (Bahuguna et al., 2022). In Geraldton and Northampton on the northern tip of the West Australian wheatbelt, the mean minimum temperatures between October and December (i.e. during reproductive development to harvest) range from 12 °C to 17 °C. Similar temperatures, in controlled-environment settings, enable dissection of the physiological mechanism of tolerance and application of high-throughput phenotyping platforms to screen for stress tolerance; for example, see Wang et al. (2022) with control nights of 21 °C and high nights of 28 °C. Experiment II was a repetition of Experiment I except plants with mature (pre-existing) flag leaves which had grown under 15 °C nights were exposed to a shorter period of warming (5–7 nights at 20 °C and 25 °C environments) at anthesis. As such Experiment I focused on newly developed leaves and Experiment II on pre-existing leaves. For both experiments, all measurements were taken when plants were at the same developmental stage—anthesis (i.e. between Zadok scale ZS60 and ZS69; Zadoks et al., 1974).

Plant materials, management, and temperature treatments

Four wheat (T. aestivum L.) genotypes were used for this study: Mace (pedigree Wyalkatchem/Stylet/Wyalkatchem), a well-adapted, commercial, Australian cultivar; ACIAR09PBI C38-150C-DH9 (pedigree PBW343+L24+LR28/LANG; henceforth 1704), a heat-susceptible genotype; ACIAR09PBI C27-0C-0N-3N (pedigree DBW16/ANNUELLO; henceforth 1898), also a heat-susceptible genotype; and 8:ZW11 [pedigree D67.2/P66.270//AE.SQUARROSA(320)/3/CUNNINGHAM/4/VORB; henceforth 2254], a heat-tolerant elite genotype. Mace was a benchmark variety for yield in Western Australia where it was widely grown. It accounted for 66% of plantings between 2015 and 2016 (Zaicou-Kunesch et al., 2017), although this reduced to 31% by 2018. Mace can cope with high temperatures (Bokshi et al., 2022). The two heat-susceptible genotypes (1704 and 1898) exhibit low PSII thermal stability (i.e. low basal Tcrit) when grown under non-heat-stressed conditions but also the capacity to acclimate Tcrit to warmer growth regimes in field conditions (Posch et al., 2022b). Genotype 2254 was developed by the International Maize and Wheat Improvement Centre (CIMMYT) and, like Mace, it is considered heat tolerant (Posch et al., 2022a). Unpublished yield data from field experiments conducted under heat stress conditions in Ciudad Obregon, Mexico and Narrabri, Australia by Professor Richard Trethowan of the University of Sydney, Australia, support the heat-tolerant classification of genotype 2254. These were chosen for their varied yield, and agronomic and physiological performance under heat stress conditions in fields across wheat-growing regions of Australia.

Seeds were germinated on moist filter papers in Petri dishes. One-week-old seedlings were sown into 6 litre plastic pots (one seedling per pot) filled with Martins mix (Martins Fertilizers Ltd, Yass, NSW Australia). The Martins mix was enriched with 4 g l–1 Osmocote® OSEX34 EXACT slow‐release fertilizer (Scotts Australia, Bella Vista, NSW, Australia) and treated at 63 °C for 1 h prior to filling pots. Potted seedlings were transferred into temperature-controlled growth chambers (Thermoline, Wetherhill Park, Australia) at the Controlled Environment Facilities of the Research School of Biology, The Australian National University, Canberra, Australia (ANU). Genotypes were arranged randomly in each chamber. Growth chambers were maintained at day/night temperatures of 26/15, 26/20. and 26/25 °C either throughout plant growth (lasting 9–13 weeks; Experiment I) or only at anthesis (lasting 5–7 d; Experiment II). For the latter conditions, day/night temperature prior to anthesis was maintained at 26/15 °C. In all growth chambers, relative humidity varied from 30% during the light period to 70% during the dark, and [CO2] was at ambient ~400 μmol mol–1 (38.4 Pa, considering a mean atmospheric pressure of 96 kPa at ANU). Lighting was supplied by 1000 W metal halide lamps (Multi-Vapor®; GE Lighting, Derrimut, Australia) producing photosynthetically active radiation at a plant height of 720–750 µmol m–2 s–1. A 12 h photothermal regime was maintained throughout plant growth. General plant management followed the established protocol of the Controlled Environment Facilities at ANU. Eight replicate plants of each genotype were assigned to each night temperature treatment.

Determination of flag leaf photosynthetic heat tolerance

T crit was estimated according to the method of Schreiber and Berry (1977) and recently applied by Zhu et al. (2018), Arnold et al. (2021), and Coast et al. (2021b). A detailed description of the method is given in Coast et al. (2021b); briefly, discs excised during the day from the middle section of detached dark-adapted leaves were exposed to a temperature ramp at a constant rate of 1 °C min−1 from 20 °C to 65 °C with simultaneous continuous measurement of F0 taken. Tcrit was calculated as the intersection point of two regression lines extrapolated from the flat and steep portion of the F0–temperature response curve. Tcrit was determined for all four genotypes in Experiment I, and for two genotypes (Mace and 1704) in Experiment II.

Gas exchange measurements

Plants in growth chambers were moved into a temperature-controlled cabinet (Thermoline Model‐1175-SD-1SL, Thermoline Scientific, Smithfield, NSW, Australia) for all gas exchange measurements. Gas exchange measurements were conducted on intact flag leaves of the main tiller identified as the first tiller to reach anthesis. Five LI-COR portable photosynthesis systems (LI-6400XT, LI-COR Inc., Lincoln, NE, USA) were used for gas exchange measurements. The LI-COR units were fitted with 6 cm2 leaf chambers with a red–blue light source (6400-18 RGB Light Source, LI-COR). Leaves were exposed to saturating irradiance of 1500 μmol photons m−2 s−1 within the LI-COR leaf chamber, with both the LI-COR leaf chamber/block and the whole plant placed within the temperature-controlled cabinet. The LI-COR leaf chamber was initially set to 20 °C, reference line atmospheric [CO2] of 400 ppm, a flow rate of 500 μmol s−1, and relative humidity maintained between 40% and 75%. After flag leaves had been exposed to these conditions in the leaf chamber for at least 5 min and following equilibrium (stable readings for at least 1 min), An was determined. Thereafter photosynthetic [CO2] response curves (A:Ci curves) were generated, at a constant irradiance of 1500 μmol photons m−2 s−1, by varying the [CO2] inside the LI-COR leaf chambers as follows: 30, 50, 100, 150, 250, 400, 400, 600, 800, 1000, 1200, 1400, and 400 μmol mol−1. The A:Ci curves were repeated with the leaves exposed to measurement temperatures of 25, 30, 35, 40, and 50 °C. The temperature setting of the cabinet was adjusted to enable the LI-COR leaf chamber/block to achieve the desired measurement temperatures. Plants were kept well irrigated throughout the measurement period to avoid water stress. In both Experiment I and II, four replicate plants per night temperature treatment were used. All gas measurements were taken between 08.00 h and 17.00 h during weeks 9–13 for Experiment I or days 5–7 for Experiment II. The ranges (9–13 weeks, and 5–7 d) in period reflect differences in time to anthesis of the different genotypes.

Modelling photosynthetic capacity

Model parameters for each growth and measurement temperature were estimated following the FvCB model and using the Plantecowrap package (Stinziano et al., 2018) in the R computing environment (R Development Core Team, 2021). Plantecowrap eliminates potential bias associated with manually determining the concentration of intercellular CO2 (Ci) where one limitation transitions to another, as is the case with the commonly used Sharkey Excel spreadsheet (Sharkey et al., 2007), and allows for species-specific kinetic parameters to be pre-defined. The kinetic parameters used in modelling photosynthetic capacity for wheat were: mesophyll conductance at 25 °C (gm=5.5 μmol m−2 s−1 Pa–1); activation energy of mesophyll conductance (Ea=47.65 kJ mol–1); apparent Michaelis–Menten constant for Rubisco carboxylation in 21% oxygen (Kair=772 µmol mol–1); activation energy of Kc (93.72 kJ mol–1); photorespiratory CO2 compensation point or Gamma star at 25 °C (Γ*=37.74 µmol mol–1, equivalent to µbar bar–1); and Gamma star activation energy (24.42 kJ mol–1). The temperature responses of Vcmax, J1500, and TPU were modelled using non-linear least squares fit of the Arrhenius temperature response function accounting for deactivation (Medlyn et al., 2002; Kattge and Knorr, 2007). The deactivation energy (Ed) was assumed to be 200 kJ mol–1; the activation energy (Ea) and entropy factor (ΔS) were estimated from iterative fits of the model.

Gas chromatography–mass spectrometry metabolite analysis

All leaf samples used for metabolite analysis were collected during the day (between 08.00 h and 17.00 h) and within an hour of completing gas exchange measurements. Metabolite extraction was conducted using a GC-MS procedure described in Rashid et al. (2021). Derivatization was performed using the MPS2 XL-Twister autosampler (Gerstel GmbH & Co. KG, Mülheim an der Ruhr, Germany), and metabolite samples were analysed on an Agilent GC/MSD system composed of an Agilent GC 6890N gas chromatograph (Agilent Technologies) fitted with a 7683B Automatic Liquid Sampler (Agilent Technologies) and 5975B Inert MSD quadrupole MS detector (Agilent Technologies). The gas chromatograph was fitted with a 0.25 mm (i.d.), 0.25 μm film thickness, 30 m Agilent FactorFour VF-5ms capillary column with a 10 m integrated guard column (Agilent Technologies). Raw GC-MS data were converted using the GC/MS Translator (ver 1.0, Agilent Technologies, Inc., Santa Clara, CA, USA) then processed (peak detection, retention time alignment, and relative quantitation) with MS-DIAL (ver 4.60; Tsugawa et al., 2020). Metabolites were identified by comparing mass spectral features against available spectral libraries from the Golm Metabolome Database (Kopka et al., 2004). Metabolite abundance values were normalized against the averaged signal of an internal standard ([13C5]valine) and the sample fresh mass, followed by weighting against the average measured signal across all samples for each compound.

Statistical analysis

Data preparation, analysis, and visualization were performed with R (R Development Core Team, 2021) using the packages tidyverse (Wickham et al., 2019), tidyr (Wickham, 2021), dplyr (Wickham et al., 2021), ggplot2 (Wickham, 2016), FactoMineR (et al., 2008), factoextra (Kassambara and Mundt, 2020), ade4 (Chessel et al., 2004; Dray and Dufour, 2007; Dray et al., 2007; Bougeard and Dray, 2018; Thioulouse et al., 2018), nls2 (Grothendieck, 2013), minpack.lm (Elzhov et al., 2016), vegan (Oksanen et al., 2020), lattice (Sarkar, 2008), latticeExtra (Sarkar and Andrews, 2022), car (Fox and Weisberg, 2019), AICcmodavg (Mazerolle, 2020), and agricolae (de Mendiburu, 2021).

Experiments I and II were analysed separately. Gas exchange data were checked for outliers prior to analysis. Differences in treatment means for Tcrit, gas exchange parameters (Amax and gs), and modelled photosynthetic capacities (Vcmax and J1500 at a standardized temperature of 25 °C; Topt of An, Vcmax, J1500, and TPU; and Vcmax, J1500, and TPU at Topt) were tested using ANOVA and treatments separated using Fisher’s LSD or Tukey’s HSD. Comparative regressions were used to test the effects of night temperature and genotype on the temperature response curves of An, Vc, J, and TPU. The regression models tested fits of a common line (the simplest model) and separate lines (the advanced model) for the genotype and night temperature. The simplest model assumed a common intercept and curvature, whereas the advanced model assumed differences in intercept or curvature, or both. F-tests were used to select the model with the best fit. The Amax (i.e. An at Topt) and Topt of An were derived from equations of the second-order polynomials that best described the instantaneous An–temperature relationships. Metabolite abundance data were analysed using principal component analysis (PCA) and a non-parametric, permutation tests-based, multivariate ANOVA (Anderson, 2001). This form of analysis generates test statistics analogous to Fisher’s F-ratio, and P-values are obtained using permutations.

Results

T crit of newly developed leaves (i.e. leaves that developed entirely under conditions where the prevailing night temperature was elevated) varied between genotypes (P<0.001) but did not significantly respond to night warming (Fig. 1A). Across night temperatures, mean Tcrit of the heat-susceptible genotype 1704 (45.7 °C) was higher (P<0.05) than the Tcrit of Mace (43.8 °C) and the elite genotype 2254 (44.6 °C), but not of genotype 1898 (45.2 °C, P=0.503). In pre-existing leaves (i.e. leaves that had developed under controlled-environment conditions and then exposed to 5–7 consecutive warm nights), responses of both genotypes to night warming differed: Tcrit of Mace increased with warm nights whereas Tcrit of 1704 did not significantly change, resulting in a significant genotype by night temperature interaction (P=0.008, Fig. 1B). Across night temperatures, Mace and 1704 differed in Tcrit (P=0.015, for main effect of genotype). However, the effect of night temperature alone was not significant (P=0.107).

Fig. 1.

Fig. 1.

Photosynthetic high temperature tolerance (Tcrit) of newly developed leaves of four wheat genotypes (top panel) and pre-existing leaves of two wheat genotypes (bottom panel) at night temperatures of 15, 20, and 25 °C. The four genotypes were the heat-susceptible 1704 and 1898, and the heat-tolerant Mace and 2254. Means are of 3–8 plants.

When measured at ambient CO2, instantaneous temperature responses of light-saturated rates of An in newly developed wheat leaves responded to night warming (Fig. 2A; Table 1). The responses were characterized by maintaining or increasing the maximum rate of An (i.e. Amax at Topt) or the optimum temperature of An (Topt of An). The largest increase with night warming from 15 °C to 25 °C in Amax at Topt and Topt of An were in genotype 2254 and 1898, respectively. However, for genotype 1704, estimates of Amax at Topt and Topt of An for plants grown at 20 °C and 25 °C nights could not be determined due to the linear declines in An within the limited measurement temperature range studied. For genotype 1704 alone, comparison of regressions for the different night temperatures showed variation [F=65.07(4, 68), P<0.001, and R2=0.79], but this was due to a lower An–temperature response at 20 °C compared with both 15 °C and 25 °C nights. Stomatal conductance (gs) also responded to night warming, and was generally curvilinear with increasing measurement temperature (Fig. 2B). There was a clear interaction effect of genotype by night temperature on gs, with lower gs for genotypes 1704 and 1898 under warmer nights, and the opposite for Mace and genotype 2254. The short-term leaf temperature response of gs and An were curvilinear. Stomatal conductance was marked by an initial decline as temperature increased from 20 °C to ~30 °C, followed by an uptick at higher temperatures. There were minimal changes in photosynthesis for leaves that pre-existed when night temperatures were altered (Fig. 2C).

Fig. 2.

Fig. 2.

At ambient CO2: instantaneous temperature response curves of light-saturated net assimilation (An; top panels) and stomatal conductance (gs; bottom panels) in wheat flag leaves that developed (newly developed) at a day temperature of 20 °C and night temperatures of 15, 20, or 25 °C. Individual panels are presented for the heat-susceptible genotypes 1704 (A) and 1898 (B), and the heat-tolerant genotypes 2254 (C) and Mace (D). For pre-existing leaves that only experienced night warming after spikes emerged and flowered, the instantaneous An–temperature responses were unaffected by night warming (Supplementary Fig. S1). Lines are polynomial regression curves fitted to the data. The shaded regions indicate confidence intervals for the fitted polynomial curves. Means are of 2–4 plants.

Table 1.

Estimated optimum temperature (Topt) of light-saturated net assimilation at ambient CO2 (An) in leaves of four wheat genotypes, and An at Topt (i.e. maximum An or Amax) derived from regression models of instantaneous temperature responses of An

Night temperature Genotypesa
1704 1898 2254 Mace
Newly developed leavesb
T opt of An 15 16.0 22.6 29.9 24.6
20 --- 22.3 26.0 25.2
25 --- 26.2 24.7 11.7
A max (An at Topt) 15 30.5 26.0 25.2 28.1
20 --- 20.5 25.6 31.9
25 --- 19.3 35.1 31.6
Pre-existing leaves
T opt of An 15–25 18.6 --- --- 24.4
A max (An at Topt) 15–25 21.0 --- --- 23.4

a The heat-susceptible genotypes are 1704 and 1898, and the heat-tolerant genotypes are 2254 and Mace.

b The final models for newly developed leaves of each genotype were based on separate regression fits for the three different night temperatures, whereas for pre-existing leaves models were common regressions across the night temperature. --- The best fit was linear, Topt could not be determined within the limits of temperature range tested.

V cmax of newly developed leaves increased exponentially with rising leaf temperatures before peaking then declining at measurement temperatures above 36–47 °C for Vcmax and 29–34 °C for J1500 (Fig. 3; Table 2), which enabled a classical Arrhenius temperature response with a deactivation component to be fit. Topt of Vcmax was unresponsive to night warming (P=0.112, Table 2) but modelled rates of Vcmax at Topt increased significantly, between 24% and 43% (Fig. 3; Table 2). Below leaf temperatures of 36 °C, Vcmax and its response to temperature were similar between the genotypes and not affected by night temperature. Thus, the impact of night warming was most evident as an increase in Rubisco carboxylation capacity of leaves developed under the warmest nights.

Fig. 3.

Fig. 3.

Temperature response curves of the maximum CO2 carboxylation capacity (Vcmax), photosynthetic electron transport capacity (J1500), and triphosphate utilization (TPU) of leaves that developed (newly developed) at a common day temperature of 20 °C and night temperatures of 15, 20, or 25 °C for four wheat genotypes. The genotype name/ID are indicated in the panels for both the heat-susceptible genotypes 1704 (A) and 1898 (B), and the heat-tolerant genotypes 2254 (C) and Mace (D). Vcmax was iteratively fit with an Arrhenius equation with parameters given in Table 2. J1500 and TPU were fit with quadratic functions. Plots for pre-existing leaves of Mace and 1704 are presented in Supplementary Figure S2. Means are of 3–4 plants. Error bars represent the standard error of the mean.

Table 2.

Mean estimates (±SD) of the optimum temperature (Topt) and maximum capacity of Vcmax, J1500, and TPU in newly developed leaves of four wheat genotypes at three different night temperatures

Genotypea Night temperature (°C) T opt of Vcmax (°C) V cmax at Topt (μmol m–2 s–1) T opt of J1500
(°C)
J 1500 at Topt (μmol electrons m–2 s–1) T opt of TPU
(°C)
TPU at Topt (μmol m-2 s-1)
1704
15 38.1 ± 7.1 310 ± 32 29.4 ± 1.8 222 ± 8 29.6 ± 1.7 15 ± 2
20 38.7 ± 0.2 217 ± 2 31.5 160 28.9 11
25 41.6 ± 0.6 407 ± 28 32.8 ± 0.2 192 ± 17 32.6 ± 4.0 12 ± 4
1898
15 37.3 ± 3.4 242 ± 26 39.9 ± 1.6 195 ± 15 29.5 ± 2.0 13 ± 2
20 --- --- 29.2 ± 1.4 179 ± 35 29.0 ± 0.9 12 ± 1
25 46.1 ± 4.8 346 ± 94 33.7 ± 6.1 181 ± 14 32.5 ± 7.0 11 ± 7
2254
15 39.3 ± 5.1 288 ± 32 29.2 ± 0.1 181 ± 37 28.5 ± 0.1 12 ± 3
20 46.6 436 ± 2 34.1 ± 0.3 208 ± 20 32.6 ± 2.6 13 ± 2
25 38.9 ± 2.3 498 ± 28 31.6 ± 1.5 259 ± 6 30.3 ± 2.3 16 ± 2
Mace
15 36.2 ± 6.2 326 ± 58 29.8 ± 3.4 213 ± 49 28.5 ± 4 13 ± 4
20 39.1 ± 5.6 389 ± 33 30.8 ± 0.8 251 ± 23 30.1 ± 0.9 16 ± 1
25 46.0 ± 9.8 569 ± 176 32.3 ± 6.5 225 ± 31 30.7 ± 6.9 14 ± 7
Levels of significance (P-values)
Genotype (G) 0.963 0.186 0.979 0.024 0.961 0.178
Night temperature (NT) 0.112 0.034 0.156 0.851 0.330 0.942
G x NT 0.568 0.796 0.854 0.047 0.922 0.078

a The heat-susceptible genotypes are 1704 and 1898, and the heat-tolerant genotypes are 2254 and Mace. --- The model fit could not be resolved. Highlighted in bold are P-values <0.05.

J 1500 and TPU displayed a curvilinear response to leaf temperature (Fig. 3E, F). The genotype by night temperature effect on J1500 was significant (P=0.047, Table 2). This was due in part to J1500 decreasing with night warming for genotypes 1704 and 1898 but not being detrimentally affected by night warming in Mace and genotype 2254. J1500 at Topt for Mace and 2254 increased by 5–30% with warmer nights (Table 2). By contrast, night warming reduced J1500 in 1704 and 1898 by 7–28% (Table 2). TPU was largely unresponsive to night warming and did not differ between genotypes. Like Topt of Vcmax, Topt of J1500 and TPU were not significantly changed under warmer nights. However, Topt values of photosynthetic capacity were high for Vcmax (range 37–47 °C) relative to J1500 (range 29–34 °C) and least for TPU (range 11–16 °C). As with net assimilation rates, photosynthetic capacity in terms of Vcmax, J1500, and TPU were not influenced by night warming in pre-existing leaves of both genotypes (P>0.05, Supplementary Fig. S1; Supplementary Table S1).

To determine whether light-saturated An becomes RuBP limited in newly developed leaves, as suggested by the decline in J1500, An was plotted against Ci for the different night temperatures. The curvilinear responses were fitted using the modelled carboxylation and RuBP regeneration limitation rates (Fig. 4). At Ci corresponding to an ambient atmospheric CO2 concentration of 400 µmol mol–1, light-saturated An was reduced in 1704 and 1898 in response to a greater RuBP regeneration limitation at 20 °C and 25 °C nights (see black outlined boxes in Fig. 4). Genotypes 2254 and Mace either maintained or increased their RuBP regeneration capacity with night warming and consequently An at 400 µmol mol–1 across night treatments.

Fig. 4.

Fig. 4.

ACi curves and corresponding C3 photosynthesis model fits for newly developed leaves of wheat genotypes 1704 (A, heat-susceptible), 1898 (B, heat-susceptible), 2254 (C, heat-tolerant), and Mace (D, heat-tolerant). Plants were treated to a common day temperature of 20 °C and night temperatures of 15 (purple lines and shapes), 20 (green lines and shapes), or 25 °C (yellow lines and shapes). Solid, dashed, and dotted curves represent predicted carboxylation-limited An rates, RuBP regeneration-limited An rates, and triphosphate utilization- (TPU) limited An rates, respectively. The black outlined box encapsulates the values corresponding to ambient CO2 concentration of 400 μmol mol–1. Values are the means of pooled measurements at leaf temperatures of 25 °C and 30 °C and model parameters standardized to 28 °C. Means are of 4–6 plants.

Across the four genotypes and night temperatures of Experiment I, a total of 162 metabolites were identified in newly developed leaves. PCA determined relationships among individual metabolites and across treatments. Metabolites were assigned to classes as listed in Fig. 5 (metabolite relative abundance and classification are presented in Supplementary Table S2). The response of metabolites generally followed a similar pattern for a particular class, and often a class responded independently of other classes. Most notably, organic acids and amino acids clustered in the right half of the PCA (positive Dim1), while many carbohydrates clustered in the opposite left half. Thus, organic and amino acid abundance were in general positively related to one another and negatively correlated with many carbohydrates. In terms of night treatment, there was a shift to a more distinct separation of metabolite abundance in genotypes that were more physiologically responsive and tolerant to the warmer nights (Fig. 5). For example, the most sensitive and non-responsive genotype 1704 had little separation of metabolite profiles among night temperatures—evident in overlapping night temperature-dependent metabolite distribution clouds in Fig. 5. By contrast, the two most heat-responsive and tolerant genotypes (Mace and 2254) had a distinct special separation of the 25 °C night treatment metabolites, represented by special separation at negative Dim1 and Dim2 values (i.e. the lower left corner of PCA graphs). Genotype 1898 had intermediate separation of metabolite profiles in response to night warming, shifting to negative Dim1 values. Interestingly, many of the same metabolites positively correlated with warmer nights for the three genotypes (1898, 2254, and Mace) that responded to warmer nights (Fig. 6). Of these metabolites that responded positively, most were monosaccharides, namely 3,6-anhydro-d-hexose, fructose, mannose, mannose-6-phosphate, and tagatose (metabolites 14, 55, 113, 114, and 149, respectively). The other metabolites that consistently responded to the warmest night of 25 °C were the saturated 18:0 and 16:0 fatty acids octadecanoic acid and palmitic acid (metabolites 124 and 129).

Fig. 5.

Fig. 5.

Principal component analysis (PCA) decomposition of metabolite responses in newly developed leaves of wheat genotypes to three different night temperature. The genotype name/ID are indicated in the panels for both the heat-susceptible genotypes 1704 and 1898, and the heat-tolerant genotypes 2254 and Mace. Metabolites are colour-coded by functional class. The direction and length of vector arrows provide information on the contribution each metabolite makes to dimensions 1 (Dim 1) and 2 (Dim 2), and the inter-relatedness of metabolites among themselves (opposing directions represent strong negative correlations while arrows of similar direction represent positive correlations), as well as in relation to night temperatures (the latter represented by blue, green, and red shaded areas). Arrow numbers correspond to individual metabolites which are listed in Supplementary Table S1, along with their relative abundance.

Fig. 6.

Fig. 6.

Individual metabolites that consistently responded positively to the warmest night of 25 °C in the genotypes 1898 (open circles; heat-susceptible), 2254 (open squares; heat-tolerant), and Mace (open triangles; heat tolerant). Metabolites presented are monosaccharides (A–D) and fatty acids (E, F). The genotype 1704 was excluded from this analysis as its metabolite profile showed no significant response to night temperature, evident in the PCA plots presented in Fig. 5. The metabolite numbers refer to their corresponding labels in the PCA plots.

Discussion

In this study, we showed that wheat photosynthesis and acclimation capacity are directly affected by night warming, with varying genotypic responses and differences between newly developed and pre-existing leaves. For most genotypes, metabolites for monosaccharides and saturated fatty acids had the strongest positive correlation to warmer nights. Osmotic regulation and changes in fatty acid saturation with warming, which is consistent with changing membrane structural integrity (Rudolph and Goins, 1991; Harishchandra et al., 2010; Marček et al., 2019; Siddiqui et al., 2020), may be linked to the heat tolerance potential of wheat to night warming. Difference between newly developed and pre-existing leaves were consistent with previous physiological reports of plants demonstrating a far greater acclimation potential in tissue that developed under a particular thermal regime rather than simply experiencing thermal change post-development (Armstrong et al., 2006; Campbell et al., 2007; Rashid et al., 2020). However, we note that our results are based on a small number of genotypes, with a focus on photosynthesis-related traits measured at a single growth stage (anthesis) in controlled-environment conditions. Thus, possible conclusions about mechanisms of high night temperature tolerance are limited and deserve further attention in future studies.

Heat-tolerant wheat genotypes can sustain Vcmax at higher leaf temperatures when grown at warmer nights

Two heat-tolerant genotypes (Mace, a locally adapted commercial cultivar; and 2254, an elite heat-tolerant genotype) were able to either maintain or increase CO2 fixing capacity at hotter leaf temperatures in response to warmer nights (Fig. 3). This was most evident in the significantly greater Vcmax reached at its Topt when grown under warmer nights (Table 2; Fig. 3). An increase in photosynthetic capacity through an increase in Vcmax and its Topt, in response to both day and night warming, is evident across a range of plant species (Kattge and Knorr, 2007; Smith and Dukes, 2017). The fact that we observed these acclimation responses in wheat genotypes due to increases in night temperature alone is interesting considering that the night temperature was not expected to have a direct influence on photosynthesis due to the temporal separation of the night warming from photosynthesis in C3 species. Night-time respiration is more likely to be a major contributor to acclimation than photosynthesis. Although we have focused on photosynthesis and not respiration, we expect that night warming will alter the ratio of dark respiration to An, similar to previous studies (Turnbull et al., 2002; Bahuguna et al., 2017; Impa et al., 2019).

The stable to higher Vcmax at hotter leaf temperature in 2254 and Mace with warming nights was probably a result of more active Rubisco rather than a change in its abundance or kinetics (Scafaro et al., 2023). It is well established, including in wheat, that a proportion of Rubisco is inactive, and this inactivity increases with leaf temperature (Law and Crafts-Brandner, 1999; Sharwood et al., 2016; Perdomo et al., 2017). The decline in Vcmax at the hotter leaf temperatures is consistent with Rubisco becoming inactive. Declines in Rubisco activity with heat are not due to Rubisco per se which is a thermally stable protein (Salvucci et al., 2001), but rather due to the heat-labile nature of Rubisco activase (Rca), its accessory protein (Salvucci and Crafts-Brandner, 2004). The stable to higher Vcmax of the heat-tolerant genotypes at hot leaf temperatures—and only when grown at warmer nights—suggests that night warming is inducing changes in Rca and its control over Rubisco. Recently, a thermally stable isoform of Rca (Rca1β) was identified in wheat, and its expression and abundance increase with exposure of wheat to heat stress (Scafaro et al., 2019; Degen et al., 2021; Perdomo et al., 2021). Night warming-induced expression of Rca1β in 2254 and Mace but not in 1704 and 1898 would confirm this hypothesis.

The shift in Vcmax with night temperature could potentially be due to temperature-dependent shifts in the kinetic properties of Rubisco. However, there are limited indications that Rubisco kinetics may alter in response to growth conditions. One example is spinach (Spinacia oleracea), when grown at the hotter day/night temperature of 30/25 °C compared with 15/10 °C (Yamori et al., 2006). A small but significant increase in the thermal stability of spinach Rubisco implied acclimation-induced alterations in Rubisco functionality. A wider study, which included wheat, demonstrated that Rubisco properties such as its CO2 substrate affinity can differ depending on growth temperature (Orr et al., 2016). More studies are needed to conclusively determine if Rubisco isoforms and associated kinetics within a species can shift in response to growth temperature. What is most unlikely is that night warming increased Rubisco abundance, as this would manifest as an increase in Vcmax across all measured leaf temperatures, but night warming did not influence Vcmax at leaf temperatures below 36 °C (Fig. 3). Furthermore, growth of wheat under both hotter days and nights does not appear to change Rubisco abundance (Perdomo et al., 2017).

CO2 conductance and electron transport limit photosynthesis in the heat-susceptible genotypes with night warming

Newly developed leaves of the genotype 1898 and the heat-susceptible genotype 1704 (known to be susceptible in terms of yield and growth in the field under hotter growth conditions) exhibited reduced photosynthesis rates when night temperature was warmer than 15 °C. The detrimental effects of night warming on the photosynthetic activity could be attributed to a decline in CO2 substrate availability, electron transport capacity, and TPU under the warmer nights (Figs 24). Declines in gs with night warming are associated with a decline in the Ci/Ca ratio. A decline in Ci/Ca (as a result of decreased gs) without an increase in An is a strong indication of water stress and CO2 substrate limitations on photosynthetic rates (Morison and Gifford, 1983; Condon et al., 2004). Thus, it is likely that these wheat genotypes had a sensitive response to perceived limitations in water availability triggered by night warming, or adopted reduced gs by increasing abscisic acid in direct response to heat stress (Rodriguez and Davies, 1982; Li et al., 2020). It may also have been due to the link between temperature or VPD and stomatal conductance. For the well-adapted Mace and heat-tolerant elite genotype 2254, the strong coupling of gs with An (Lawson et al., 2011), and other mechanisms linked to water status (e.g. increase in xylem and mesophyll hydraulic conductance; Urban et al., 2017a), would explain the parallel increases in gs and An at higher nocturnal temperature. The assumption of a fixed gsAn relationship is central to many models of stomatal control of photosynthesis at different scales: leaf, plant, ecosystem, and global circulation models (Farquhar and Wong, 1984; Leuning 1995; Buckley et al., 2003; Verhoef and Egea, 2014). However, this gsAn relationship can be decoupled under extreme conditions such as heatwaves (Ameye et al., 2012; von Caemmerer and Evans, 2015) or at extremely high leaf temperatures (Urban et al., 2017b).

Night warming above 15 °C reduced the photosynthetic electron capacity and/or the Topt of J1500 of genotypes 1704 and 1898 (Fig. 3; Table 2). This reduction in J1500 led to a modelled RuBP regeneration limitation in An at current ambient CO2 concentrations (Fig. 4). Despite the reductions in J1500 and/or the Topt of J1500 with night warming, PSII was found to be thermally stable (higher Tcrit) than the heat-tolerant Mace and 2254 (Fig. 1A). The mismatch between Tcrit and J1500 demonstrates that the upper temperature limit of electron transport is not necessarily aligned with susceptibility of electron transport rates under the prevailing growth temperature. In other words, photosynthesis of 1898 and 1704 may have an ability to withstand a more severe heat shock despite being more susceptible to loss of function under milder but more sustained heating. Genotype 2254 was able to increase J1500 and had relatively larger increases in Tcrit with night warming than 1704 and 1898, possibly meaning that the extent of change in Tcrit is more informative than the upper limit of Tcrit when assessing the potential for electron capacity to adjust to temperature. Pre-existing leaves of Mace, but not 1704, responded strongly to warmer nights. This suggests a greater role for older leaves in heat stress resilience, even though these leaves are assumed to have less capacity to acclimate to high temperatures.

In contrast to Vcmax, the response of J1500 and TPU to leaf measuring temperatures did not follow an expected exponential rise that could be fit with an Arrhenius equation (Fig. 3). A similar finding has previously been observed in wheat for J (Silva-Perez et al., 2017). This is likely to be linked to Vcmax being driven by the kinetics of Rubisco enzyme and its exponential response to temperature (apart from the aforementioned decline at hotter leaf temperatures due to Rca loss of function). J1500 is set by more complex protein- and membrane-dependent interactions, some of which may not respond as dynamically to short-term rises in leaf temperature. TPU, on the other hand, largely comes into play at high CO2 levels, beyond the ambient conditions under which these experiments were conducted.

Monosaccharide and saturated fatty acid accumulation are key responses of metabolite to warm nights

The most striking change in metabolites was a build-up of monosaccharide soluble sugars in newly developed leaves of the more heat-tolerant genotypes when grown at warmer nights. The build-up of monosaccharides may be related to maintenance of the osmotic potential of cells (Marček et al., 2019) and regulation of stomatal conductance, which is also linked to photosynthesis (Flütsch et al., 2020). Soluble sugars are potent contributors to cell osmotic potential and regulation (Moinuddin et al., 2005; Blum, 2017). It is noted that osmolyte adjustments contribute to drought and heat stress avoidance in wheat (Blum and Pnuel, 1990). Specific to night warming, the disaccharide trehalose accumulated in response to night warming in wheat spike tissue (Impa et al., 2019). Thus, the accumulation of soluble sugars in the heat-responsive genotypes but not in 1704 under warmer nights may have contributed to greater osmotic protection during warmer nights. The assumed decline in osmotic potential of the heat-responsive 2254 and Mace leaves may have contributed to its increase in gs with warmer nights, as opposed to 1704 which had a reduction in gs. Solutes, including soluble sugars, have also been shown to provide protection to membrane structural integrity during stress through greater hydration of the lipid surfaces (Rudolph and Goins, 1991; Harishchandra et al., 2010). The higher abundance of simple soluble carbohydrates together with night warming may have contributed to the maintenance of electron transport (maintained and even increased J1500) and ability to acclimate Tcrit to a greater extent than 1704. In agreement with our postulation that osmolytes are protecting J1500 in 2254 and Mace but not 1704, supplementing Populus tremula L. leaves with increasing amounts of the monosaccharide sucrose substantially reduced leaf osmotic potential and led to an increase in stability of photosynthetic electron transport at higher Tleaf (Hüve et al., 2006). That we observed a consistent increase in the saturated fatty acids octadecanoic acid and palmitic acid in the more photosynthetically heat-tolerant genotypes with night warming is a further indication of membranes adjusting to be more stable in response to warmer temperatures. Saturation of fatty acids in membranes is known to be a key heat tolerance response of plants (Larkindale et al., 2005; Zhu et al., 2018). Our results suggest that soluble sugar and saturated fatty acid contents in wheat leaves are key metabolites linking night warming and alterations in day processes. The connection between night temperature, carbohydrate and saturated fatty acid accumulation, and CO2 conductance and electron transport capacity needs further exploration. Having identified a specific heat tolerance metabolic profile, high-throughput metabolic approaches such as hyperspectral imaging could be used to build models to predict this metabolic profile as has been achieved for amino acids in maize (Shu et al., 2022). These high-throughput measures could then be deployed by wheat populations through both ground and aerial sensing applications to determine heat tolerance.

Conclusions

Night warming has significant implications on the photosynthetic performance of wheat leaves that have developed under the prevailing night temperature. The heat-tolerant genotype (2254 and Mace) had higher light-saturated An when grown at warmer nights, in contrast to the other genotypes (1704 and 1898) which had reduced An. These night temperature-dependent differences in An between genotypes of wheat could in part be attributed to how night warming influenced photosynthetic capacity and CO2 conductance. The thermally tolerant genotypes had a more stable Vcmax at higher leaf temperatures when grown at warmer nights, which is likely to be due to night-dependent alteration to the temperature sensitivity of the activation state of Rubisco. The thermally susceptible genotypes had reduced gs and J1500 when grown at warmer nights, which accounts for the corresponding reduction in An. The stability of electron transport and CO2 conductance in the heat-tolerant genotypes with night warming may be linked to the greater accumulation of monosaccharides and saturated fatty acids in its leaves, balancing osmotic pressure alterations and cell membrane integrity on exposure to warm nights. The lack of a night warming response in pre-existing leaves, as well as differences between the two wheat genotypes, demonstrates the divergence of strategies needed to improve wheat performance under future climate trends. Improving CO2 conductance and photosynthetic capacity may be beneficial in some genotypes and potentially achieved through altering monosaccharide and saturated fatty acid contents of cells. In other genotypes with no detrimental impact of night warming on photosynthetic performance (exemplified by the heat-tolerant 2254 and Mace), it might be best to focus on sink tissue limitations and other physiological processes detrimentally affected by heat and not photosynthetic capacity. An important physiological process that clearly deserves investigating is mitochondrial respiration in the dark.

Supplementary data

The following supplementary data are available at JXB online.

Fig. S1. Instantaneous temperature response curves of light-saturated net assimilation (An) and stomatal conductance (gs) in pre-existing flag leaves of four wheat genotypes at ambient CO2.

Fig. S2. Temperature response curves of the maximum CO2 carboxylation capacity (Vcmax), photosynthetic electron transport capacity (J1500), and triphosphate utilization (TPU) of pre-existing wheat leaves at night temperatures of 15, 20, or 25 °C.

Table S1. Mean estimates (±SD) of the optimum temperature (Topt) and maximum capacity of Vcmax, J1500, and TPU in pre-existing leaves of two wheat genotypes at three different night temperatures.

Table S2. Metabolite relative abundance and classification.

erad437_suppl_Supplementary_Figures_S1-S2_Tables_S1
erad437_suppl_Supplementary_Tables_S2

Acknowledgements

We acknowledge and celebrate the First Australians on whose traditional land this research was undertaken, and pay our respect to their elders past, present, and emerging. We thank the ANU Research School of Biology Plant Services team, especially Christine Larsen, Jenny Rath, Gavin Pritchard, and Steven Dempsey, for maintaining the plants in the controlled environments; and Australian Grain Technologies, Narrabri and Professor Richard Trethowan of the University of Sydney, Australia for providing the seeds. The GC-MS analysis for this work was performed by the Centre for Microscopy, Characterisation and Analysis (Metabolomics Australia), UWA and was supported by infrastructure funding from the Western Australian State Government in partnership with the Australian Federal Government, through the National Collaborative Research Infrastructure Strategy (NCRIS).

Contributor Information

Onoriode Coast, ARC Centre of Excellence in Plant Energy Biology, Research School of Biology, The Australian National University, Canberra, ACT 2601, Australia; Natural Resources Institute, University of Greenwich, Central Avenue, Chatham Maritime, Kent ME4 4TB, UK; School of Environmental and Rural Sciences, Faculty of Science, Agriculture, Business, and Law, University of New England, Armidale, NSW 2351, Australia.

Andrew P Scafaro, ARC Centre of Excellence in Plant Energy Biology, Research School of Biology, The Australian National University, Canberra, ACT 2601, Australia; Division of Plant Science, Research School of Biology, The Australian National University, Canberra, ACT 2601, Australia.

Helen Bramley, School of Life and Environmental Sciences, Plant Breeding Institute, Sydney Institute of Agriculture, The University of Sydney, Narrabri, NSW 2390, Australia.

Nicolas L Taylor, ARC Centre of Excellence in Plant Energy Biology, School of Molecular Sciences and Institute of Agriculture, The University of Western Australia, Crawley, WA 6009, Australia.

Owen K Atkin, ARC Centre of Excellence in Plant Energy Biology, Research School of Biology, The Australian National University, Canberra, ACT 2601, Australia; Division of Plant Science, Research School of Biology, The Australian National University, Canberra, ACT 2601, Australia.

Elizabete Carmo-Silva, Lancaster University, UK.

Author contributions

OC and OKA: conceptualization; OC and APS: data curation, formal analysis, visualization, and writing—original draft: OC, OKA, and HB: funding acquisition; OC: investigation; OC, APS, and NT: methodology; OC, APS, HB, and OKA: project administration; OKA, resources and supervision; OC, APS, HB, NT, and OKA: writing—review and editing.

Conflict of interest

No conflict of interest declared.

Funding

This work was supported by grants from the ARC Centre of Excellence in Plant Energy Biology (CE140100008), the Australian Grains Research and Development Corporation (GRDC) Postdoctoral Fellowship: ‘Photosynthetic acclimation to high temperature in wheat’ (US1904-003RTX–9177346), and project ‘A national approach to improving heat tolerance in wheat through more efficient carbon allocation’ (US00080). OC was supported by Research England’s Expanding Excellence in England (E3)-funded Food and Nutrition Security Initiative of the Natural Resources Institute.

Data availability

All primary data to support the findings of this study are openly available in the Dryad Digital Repository at doi: 10.5061/dryad.fqz612jx7 (Coast et al., 2023).

References

  1. Alexander LV, Zhang X, Peterson TC, et al. 2006. Global observed changes in daily climate extremes of temperature and precipitation. Journal of Geophysical Research: Atmospheres 111, doi: 10.1029/2005JD006290 [DOI] [Google Scholar]
  2. Ameye M, Wertin TM, Bauweraerts I, Mcguire MA, Teskey RO, Steppe K.. 2012. The effect of induced heat waves on Pinus taeda and Quercus rubra seedlings in ambient and elevated CO2 atmospheres. New Phytologist 196, 448–461. [DOI] [PubMed] [Google Scholar]
  3. Anderson MJ. 2001. A new method for non-parametric multivariate analysis of variance. Austral Ecology 26, 32–46. [Google Scholar]
  4. Armstrong AF, Logan DC, Tobin AK, O’Toole P, Atkin OK.. 2006. Heterogeneity of plant mitochondrial responses underpinning respiratory acclimation to the cold in Arabidopsis thaliana leaves. Plant, Cell & Environment 29, 940–949. [DOI] [PubMed] [Google Scholar]
  5. Arnold PA, Briceño VF, Gowland KM, Catling AA, Bravo LA, Nicotra AB.. 2021. A high-throughput method for measuring critical thermal limits of leaves by chlorophyll imaging fluorescence. Functional Plant Biology 48, 634–646. [DOI] [PubMed] [Google Scholar]
  6. Atkin OK, Tjoelker MG.. 2003. Thermal acclimation and the dynamic response of plant respiration to temperature. Trends in Plant Science 8, 343–351. [DOI] [PubMed] [Google Scholar]
  7. Bahuguna R, Solis C, Shi W, Jagadish SVK.. 2017. Post-flowering night respiration and altered sink activity account for high night temperature-induced grain yield and quality loss in rice (Oryza sativa L). Physiologia Plantarum 159, 59–73. [DOI] [PubMed] [Google Scholar]
  8. Bahuguna RN, Chaturvedi AK, Pal M, Chinnusamy V, Jagadish S, Pareek A.. 2022. Carbon dioxide responsiveness mitigates rice yield loss under high night temperature. Plant Physiology 188, 285–300. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Blum A. 2017. Osmotic adjustment is a prime drought stress adaptive engine in support of plant production. Plant, Cell & Environment 40, 4–10. [DOI] [PubMed] [Google Scholar]
  10. Blum A, Pnuel Y.. 1990. Physiological attributes associated with drought resistance of wheat cultivars in a Mediterranean environment. Australian Journal of Agricultural Research 41, 799–810. [Google Scholar]
  11. Bokshi AI, Thistlethwaite RJ, Chaplin ED, Kirii E, Trethowan RM, Tan DK.. 2022. Physiological traits for evaluating heat‐tolerance of Australian spring wheat cultivars at elevated CO2. Journal of Agronomy and Crop Science 208, 178–196. [Google Scholar]
  12. Bougeard S, Dray S.. 2018. Supervised multiblock analysis in R with the ade4 package. Journal of Statistical Software 86, 1–17. [Google Scholar]
  13. Buckley TN, Mott KA, Farquhar GD.. 2003. A hydromechanical and biochemical model of stomatal conductance. Plant, Cell & Environment 26, 1767–1785. [Google Scholar]
  14. Cai C, Li G, Di L, et al. 2020. The acclimation of leaf photosynthesis of wheat and rice to seasonal temperature changes in T-FACE environments. Global Change Biology 26, 539–556. [DOI] [PubMed] [Google Scholar]
  15. Cai C, Li G, Yang H, et al. 2018. Do all leaf photosynthesis parameters of rice acclimate to elevated CO2, elevated temperature, and their combination, in FACE environments? Global Change Biology 24, 1685–1707. [DOI] [PubMed] [Google Scholar]
  16. Campbell C, Atkinson L, Zaragoza-Castells J, Lundmark M, Atkin O, Hurry V.. 2007. Acclimation of photosynthesis and respiration is asynchronous in response to changes in temperature regardless of plant functional group. New Phytologist 176, 375–389. [DOI] [PubMed] [Google Scholar]
  17. CGIAR Advisory Services. 2020. CGIAR Research Program 2020 reviews: WHEAT. Rome: CGIAR Advisory Services Secretariat Evaluation Function; https://cas.cgiar.org/ [Google Scholar]
  18. Chessel D, Dufour AB, Thioulouse J.. 2004. The ade4 package-I-One-table methods. R news 4, 5–10. [Google Scholar]
  19. Coast O, Posch BC, Bramley H, Gaju O, Richards RA, Lu M, Ruan Y-L, Trethowan R, Atkin OK. . 2021a. Acclimation of leaf photosynthesis and respiration to warming in field-grown wheat. Plant, Cell & Environment 44, 2331–2346. [DOI] [PubMed] [Google Scholar]
  20. Coast O, Posch BC, Rognoni B, et al. 2021b. Photosynthetic heat tolerance in wheat: evidence for genotype-by-environment interactions. The Plant Journal 111, 1368–1382. [DOI] [PubMed] [Google Scholar]
  21. Coast O, Scafaro AP, Bramley H, Taylor NL, Atkin OK.. 2023. Data from: Photosynthesis in newly developed leaves of heat-tolerant wheat acclimates to long-term nocturnal warming. Dryad Digital Repository. doi: 10.5061/dryad.fqz612jx7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Condon AG, Richards RA, Rebetzke GJ, Farquhar GD.. 2004. Breeding for high water-use efficiency. Journal of Experimental Botany 55, 2447–2460. [DOI] [PubMed] [Google Scholar]
  23. Daas C, Montpied P, Hanchi B, Dreyer E.. 2008. Responses of photosynthesis to high temperatures in oak saplings assessed by chlorophyll-a fluorescence: inter-specific diversity and temperature-induced plasticity. Annals of Forest Science 65, 305–305. [Google Scholar]
  24. Degen GE, Orr DJ, Carmo-Silva E.. 2021. Heat-induced changes in the abundance of wheat Rubisco activase isoforms. New Phytologist 229, 1298–1311. [DOI] [PubMed] [Google Scholar]
  25. de Mendiburu F. 2021. agricolae: statistical procedures for agricultural research. R package version 13-5. https://cran.r-project.org/package=agricolae
  26. Downton WJS, Berry JA, Seemann JR.. 1984. Tolerance of photosynthesis to high temperature in desert plants. Plant Physiology 74, 786–790. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Dray S, Dufour A-B.. 2007. The ade4 package: implementing the duality diagram for ecologists. Journal of Statistical Software 22, 1–20. [Google Scholar]
  28. Dray S, Dufour AB, Chessel D.. 2007. The ade4 package-II: two-table and K-table methods. R news 7, 47–52. [Google Scholar]
  29. Easterling DR, Horton B, Jones PD, et al. 1997. Maximum and minimum temperature trends for the globe. Science 277, 364–367. [Google Scholar]
  30. Elzhov TV, Mullen KM, Spiess A-N, Bolker B.. 2016. minpack.lm: R interface to the Levenberg–Marquardt nonlinear least-squares algorithm found in MINPACK, plus support for bounds. R package version 1.2-1. https://cran.r-project.org/web/packages/minpack.lm/index.html [Google Scholar]
  31. Enami I, Kitamura M, Tomo T, Isokawa Y, Ohta H, Katoh S.. 1994. Is the primary cause of thermal inactivation of oxygen evolution in spinach PS II membranes release of the extrinsic 33 kDa protein or of Mn? Biochimica et Biophysica Acta 1186, 52–58. [Google Scholar]
  32. Fan Y, Tian M, Jing Q, Tian Z, Han H, Jiang D, Cao W, Dai T.. 2015. Winter night warming improves pre-anthesis crop growth and post-anthesis photosynthesis involved in grain yield of winter wheat (Triticum aestivum L). Field Crops Research 178, 100–108. [Google Scholar]
  33. Fan Y, Tian Z, Yan Y, Hu C, Abid M, Jiang D, Ma C, Huang Z, Dai T.. 2017. Winter night-warming improves post-anthesis physiological activities and sink strength in relation to grain filling in winter wheat (Triticum aestivum L). Frontiers in Plant Science 8, 992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Farquhar G, Wong S.. 1984. An empirical model of stomatal conductance. Functional Plant Biology 11, 191–210. [Google Scholar]
  35. Farquhar GD, von Caemmerer S, Berry JA.. 1980. A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta 149, 78–90. [DOI] [PubMed] [Google Scholar]
  36. Fisher T, Honsdorf N, Lilley J, Mondal S, Monasterio IO, Verhulst N.. 2022. Increase in irrigated wheat yield in north-west Mexico from 1960 to 2019: unravelling the negative relationship to minimum temperature. Field Crops Research 275, 108331. [Google Scholar]
  37. Flütsch S, Nigro A, Conci F, Fajkus J, Thalmann M, Trtílek M, Panzarová K, Santelia D.. 2020. Glucose uptake to guard cells via STP transporters provides carbon sources for stomatal opening and plant growth. EMBO Reports 21, e49719. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Fox J, Weisberg S.. 2019. An {R} companion to applied regression, 3rd edn. Thousand Oaks, CA: Sage. [Google Scholar]
  39. García GA, Dreccer MF, Miralles DJ, Serrago RA.. 2015. High night temperatures during grain number determination reduce wheat and barley grain yield: a field study. Global Change Biology 21, 4153–4164. [DOI] [PubMed] [Google Scholar]
  40. García GA, Miralles DJ, Serrago RA, Alzueta I, Huth N, Dreccer MF.. 2018. Warm nights in the Argentine pampas: modelling its impact on wheat and barley shows yield reductions. Agricultural Systems 162, 259–268. [Google Scholar]
  41. García GA, Serrago RA, Dreccer MF, Miralles DJ.. 2016. Post-anthesis warm nights reduce grain weight in field-grown wheat and barley. Field Crops Research 195, 50–59. [Google Scholar]
  42. Geange SR, Arnold PA, Catling AA, et al. 2020. The thermal tolerance of photosynthetic tissues: a global systematic review and agenda for future research. New Phytologist 229, 2497–2513. [DOI] [PubMed] [Google Scholar]
  43. Glaubitz U, Li X, Kohl KI, Van Dongen JT, Hincha DK, Zuther E.. 2014. Differential physiological responses of different rice (Oryza sativa) cultivars to elevated night temperature during vegetative growth. Functional Plant Biology 41, 437–448. [DOI] [PubMed] [Google Scholar]
  44. Grothendieck G. 2013. nls2: non-linear regression with brute force. R package version 0.2. https://cran.r-project.org/web/packages/nls2/index.html [Google Scholar]
  45. Harishchandra RK, Wulff S, Lentzen G, Neuhaus T, Galla H-J.. 2010. The effect of compatible solute ectoines on the structural organization of lipid monolayer and bilayer membranes. Biophysical Chemistry 150, 37–46. [DOI] [PubMed] [Google Scholar]
  46. Hein NT, Bheemanahalli R, Wagner D, et al. 2020. Improved cyber-physical system captured post-flowering high night temperature impact on yield and quality of field grown wheat. Scientific Reports 10, 22213. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Hikosaka K, Ishikawa K, Borjigidai A, Muller O, Onoda Y.. 2005. Temperature acclimation of photosynthesis: mechanisms involved in the changes in temperature dependence of photosynthetic rate. Journal of Experimental Botany 57, 291–302. [DOI] [PubMed] [Google Scholar]
  48. Hurry V, Igamberdiev A, Keerberg O, Pärnik T, Atkin O, Zaragoza-Castells J, Gardeström P.. 2005. Respiration in photosynthetic cells: gas exchange components, interactions with photorespiration and the operation of mitochondria in the light. In: Lambers H., Ribas-Carbo M, eds. Plant respiration. Advances in Photosynthesis and Respiration, vol 18. Dordrecht: Springer, 43–61. [Google Scholar]
  49. Hüve K, Bichele I, Tobias M, Niinemets U.. 2006. Heat sensitivity of photosynthetic electron transport varies during the day due to changes in sugars and osmotic potential. Plant, Cell & Environment 29, 212–228. [DOI] [PubMed] [Google Scholar]
  50. Impa SM, Sunoj VSJ, Krassovskaya I, Bheemanahalli R, Obata T, Jagadish SVK.. 2019. Carbon balance and source–sink metabolic changes in winter wheat exposed to high night-time temperature. Plant, Cell & Environment 42, 1233–1246. [DOI] [PubMed] [Google Scholar]
  51. Impa SM, Vennapusa AR, Bheemanahalli R, Sabela D, Boyle D, Walia H, Jagadish SVK.. 2020. High night temperature induced changes in grain starch metabolism alters starch, protein, and lipid accumulation in winter wheat. Plant, Cell & Environment 43, 431–447. [DOI] [PubMed] [Google Scholar]
  52. Kassambara A, Mundt, F.. 2020. Factoextra: extract and visualize the results of multivariate data analyses. R package version 1.0.7. https://cran.r-project.org/web/packages/factoextra/index.html [Google Scholar]
  53. Kattge J, Knorr W.. 2007. Temperature acclimation in a biochemical model of photosynthesis: a reanalysis of data from 36 species. Plant, Cell & Environment 30, 1176–1190. [DOI] [PubMed] [Google Scholar]
  54. Kopka J, Schauer N, Krueger S, et al. 2004. GMD@CSBDB: the Golm Metabolome Database. Bioinformatics 21, 1635–1638. [DOI] [PubMed] [Google Scholar]
  55. Lancaster LT, Humphreys AM.. 2020. Global variation in the thermal tolerances of plants. Proceedings of the National Academy of Sciences, USA 117, 13580–13587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Larkindale J, Mishkind M, Vierling E.. 2005. Plant responses to high temperature. In: Jenks MA., Hasegawa PM. eds. Plant abiotic stress. Oxford: Blackwell Publishing Ltd, 100–144. [Google Scholar]
  57. Law RD, Crafts-Brandner SJ.. 1999. Inhibition and acclimation of photosynthesis to heat stress is closely correlated with activation of ribulose-1,5-bisphosphate carboxylase/oxygenase. Plant Physiology 120, 173–182. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Lawson T, von Caemmerer S, Baroli I.. 2011. Photosynthesis and stomatal behaviour. In: Luttge U, ed. Progress in Botany vol. 72. Berlin, Heidelberg: Springer, 265–304. [Google Scholar]
  59. Lê S, Josse J, Husson F.. 2008. FactoMineR: an R package for multivariate analysis. Journal of Statistical Software 25, 1–18. [Google Scholar]
  60. Leuning R. 1995. A critical appraisal of a combined stomatal–photosynthesis model for C3 plants. Plant, Cell & Environment 18, 339–355. [Google Scholar]
  61. Li G, Zhang C, Zhang G, Fu W, Feng B, Chen T, Peng S, Tao L, Fu G.. 2020. Abscisic acid negatively modulates heat tolerance in rolled leaf rice by increasing leaf temperature and regulating energy homeostasis. Rice 13, 18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Lin Y-S, Medlyn BE, Ellsworth DS.. 2012. Temperature responses of leaf net photosynthesis: the role of component processes. Tree Physiology 32, 219–231. [DOI] [PubMed] [Google Scholar]
  63. Lobell DB, Ortiz-Monasterio JI, Asner GP, Matson PA, Naylor RL, Falcon WP.. 2005. Analysis of wheat yield and climatic trends in Mexico. Field Crops Research 94, 250–256. [Google Scholar]
  64. Loveys BR, Atkinson LJ, Sherlock DJ, Roberts RL, Fitter AH, Atkin OK.. 2003. Thermal acclimation of leaf and root respiration: an investigation comparing inherently fast- and slow-growing plant species. Global Change Biology 9, 895–910. [Google Scholar]
  65. Marček T, Hamow K, Végh B, Janda T, Darko E.. 2019. Metabolic response to drought in six winter wheat genotypes. PLoS One 14, e0212411. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Mazerolle MJ. 2020. AICcmodavg: model selection and multimodel inference based on (Q)AIC(c). R package version 2.3-1. https://cran.r-project.org/package=AICcmodavg
  67. Medlyn BE, Dreyer E, Ellsworth D, et al. 2002. Temperature response of parameters of a biochemically based model of photosynthesis II A review of experimental data. Plant, Cell & Environment 25, 1167–1179. [Google Scholar]
  68. Moinuddin, Fischer RA, Sayre KD, Reynolds MP.. 2005. Osmotic adjustment in wheat in relation to grain yield under water deficit environments. Agronomy Journal 97, 1062–1071. [Google Scholar]
  69. Morison JIL, Gifford RM.. 1983. Stomatal sensitivity to carbon dioxide and humidity: a comparison of two C3 and two C4 grass species. Plant Physiology 71, 789–796. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Oksanen J, Blanchet FG, Friendly M, Kindt R, Legendre P, Mcglinn D, Minchin PR, O’Hara R, Simpson G, Solymos, P.. 2020. vegan: community ecology package. R package version 2.5-7. https://cran.r-project.org/web/packages/vegan/index.html [Google Scholar]
  71. Orr DJ, Alcântara A, Kapralov MV, Andralojc PJ, Carmo-Silva E, Parry MAJ.. 2016. Surveying Rubisco diversity and temperature response to improve crop photosynthetic efficiency. Plant Physiology 172, 707–717. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. O’Sullivan OS, Heskel MA, Reich PB, et al. 2017. Thermal limits of leaf metabolism across biomes. Global Change Biology 23, 209–223. [DOI] [PubMed] [Google Scholar]
  73. Perdomo JA, Buchner P, Carmo-Silva E.. 2021. The relative abundance of wheat Rubisco activase isoforms is post-transcriptionally regulated. Photosynthesis Research 148, 47–56. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Perdomo JA, Capó-Bauçà S, Carmo-Silva E, Galmés J.. 2017. Rubisco and Rubisco activase play an important role in the biochemical limitations of photosynthesis in rice, wheat, and maize under high temperature and water deficit. Frontiers in Plant Science 8, 490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Posch BC, Hammer J, Atkin OK, Bramley H, Ruan Y-L, Trethowan R, Coast O.. 2022b. Wheat photosystem II heat tolerance responds dynamically to short- and long-term warming. Journal of Experimental Botany 73, 3268–3282. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Posch BC, Kariyawasam BC, Bramley H, Coast O, Richards RA, Reynolds MP, Trethowan R, Atkin OK.. 2019. Exploring high temperature responses of photosynthesis and respiration to improve heat tolerance in wheat. Journal of Experimental Botany 70, 5051–5069. [DOI] [PubMed] [Google Scholar]
  77. Posch BC, Zhai D, Coast O, Scafaro AP, Bramley H, Reich PB, Ruan Y-L, Trethowan R, Way DA, Atkin OK.. 2022a. Wheat respiratory O2 consumption falls with night warming alongside greater respiratory CO2 loss and reduced biomass. Journal of Experimental Botany 73, 915–926. [DOI] [PubMed] [Google Scholar]
  78. Prasad PVV, Pisipati S, Ristic Z, Bukovnik U, Fritz AK.. 2008. Impact of nighttime temperature on physiology and growth of spring wheat. Crop Science 48, 2372–2380. [Google Scholar]
  79. Rashid FAA, Crisp PA, Zhang Y, et al. 2020. Molecular and physiological responses during thermal acclimation of leaf photosynthesis and respiration in rice. Plant, Cell & Environment 43, 594–610. [DOI] [PubMed] [Google Scholar]
  80. Rashid FAA, Scafaro AP, Asao S, Fenske R, Dewar RC, Masle J, Taylor NL, Atkin OK.. 2021. Diel- and temperature-driven variation of leaf dark respiration rates and metabolite levels in rice. New Phytologist 228, 56–69. [DOI] [PubMed] [Google Scholar]
  81. Rodriguez JL, Davies WJ.. 1982. The effects of temperature and ABA on stomata of Zea mays L. Journal of Experimental Botany 33, 977–987. [Google Scholar]
  82. R Development Core Team. 2021. R: a language and environment for statistical computing. Vienna, Austria: R Foundation for Statistical Computing. [Google Scholar]
  83. Rogers A, Medlyn BE, Dukes JS, et al. 2017. A roadmap for improving the representation of photosynthesis in Earth system models. New Phytologist 213, 22–42. [DOI] [PubMed] [Google Scholar]
  84. Rudolph AS, Goins B.. 1991. The effect of hydration stress solutes on the phase behavior of hydrated dipalmitoylphosphatidylcholine. Biochimica et Biophysica Acta 1066, 90–94. [DOI] [PubMed] [Google Scholar]
  85. Russell K, Van Sanford DA.. 2020. Breeding wheat for resilience to increasing nighttime temperatures. Agronomy 10, 531. [Google Scholar]
  86. Sage RF, Kubien DS.. 2007. The temperature response of C3 and C4 photosynthesis. Plant, Cell & Environment 30, 1086–1106. [DOI] [PubMed] [Google Scholar]
  87. Salvucci ME, Crafts-Brandner SJ.. 2004. Mechanism for deactivation of Rubisco under moderate heat stress. Physiologia Plantarum 122, 513–519. [DOI] [PubMed] [Google Scholar]
  88. Salvucci ME, Osteryoung KW, Crafts-Brandner SJ, Vierling E.. 2001. Exceptional sensitivity of rubisco activase to thermal denaturation in vitro and in vivo. Plant Physiology 127, 1053–1064. [PMC free article] [PubMed] [Google Scholar]
  89. Sarkar, D. 2008. Lattice: multivariate data visualization with R. New York: Springer. [Google Scholar]
  90. Sarkar D, Andrews, F.. 2022. latticeExtra: extra graphical utilities based on lattice. R package version 0.6-30. https://CRAN.R-project.org/package=latticeExtra [Google Scholar]
  91. Scafaro AP, Bautsoens N, Den Boer B, Van Rie J, Gallé A.. 2019. A conserved sequence from heat-adapted species improves Rubisco activase thermostability in wheat. Plant Physiology 181, 43–54. [DOI] [PMC free article] [PubMed] [Google Scholar]
  92. Scafaro AP, Posch BC, Evans JR, Farquhar GD, Atkin OK.. 2023. Rubisco deactivation and chloroplast electron transport rates co-limit photosynthesis above optimal leaf temperature in terrestrial plants. Nature Communications 14, 2820. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Schaarschmidt S, Lawas LMF, Glaubitz U, Li X, Erban A, Kopka J, Jagadish SVK, Hincha DK, Zuther E.. 2020. Season affects yield and metabolic profiles of rice (Oryza sativa) under high night temperature stress in the field. International Journal of Molecular Sciences 21, 3187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Schreiber U, Berry JA.. 1977. Heat-induced changes of chlorophyll fluorescence in intact leaves correlated with damage of the photosynthetic apparatus. Planta 136, 233–238. [DOI] [PubMed] [Google Scholar]
  95. Schreiber U, Colbow K, Vidaver W.. 1975. Temperature–jump chlorophyll fluorescence induction in plants. Zeitschrift für Naturforschung C 30, 689–690. [DOI] [PubMed] [Google Scholar]
  96. Sharkey TD. 1985. Photosynthesis in intact leaves of C3 plants: physics, physiology and rate limitations. The Botanical Review 51, 53–105. [Google Scholar]
  97. Sharkey TD, Bernacchi CJ, Farquhar GD, Singsaas EL.. 2007. Fitting photosynthetic carbon dioxide response curves for C3 leaves. Plant, Cell & Environment 30, 1035–1040. [DOI] [PubMed] [Google Scholar]
  98. Sharwood RE, Sonawane BV, Ghannoum O, Whitney SM.. 2016. Improved analysis of C4 and C3 photosynthesis via refined in vitro assays of their carbon fixation biochemistry. Journal of Experimental Botany 67, 3137–3148. [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Shu M, Zhou L, Chen H, Wang X, Meng L, Ma Y.. 2022. Estimation of amino acid contents in maize leaves based on hyperspectral imaging. Frontiers in Plant Science 13, 885794. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Siddiqui H, Sami F, Hayat S.. 2020. Glucose: sweet or bitter effects in plants—a review on current and future perspective. Carbohydrate Research 487, 107884. [DOI] [PubMed] [Google Scholar]
  101. Sillmann J, Kharin VV, Zwiers FW, Zhang X, Bronaugh D.. 2013. Climate extremes indices in the CMIP5 multimodel ensemble: Part 2. Future climate projections. Journal of Geophysical Research 118, 2473–2493. [Google Scholar]
  102. Silva-Perez V, Furbank RT, Condon AG, Evans JR.. 2017. Biochemical model of C3 photosynthesis applied to wheat at different temperatures. Plant, Cell & Environment 40, 1552–1564. [DOI] [PubMed] [Google Scholar]
  103. Slot M, Rifai SW, Winter K.. 2021. Photosynthetic plasticity of a tropical tree species, Tabebuia rosea, in response to elevated temperature and [CO2]. Plant, Cell & Environment 44, 2347–2364. [DOI] [PubMed] [Google Scholar]
  104. Smith NG, Dukes JS.. 2017. Short-term acclimation to warmer temperatures accelerates leaf carbon exchange processes across plant types. Global Change Biology 23, 4840–4853. [DOI] [PubMed] [Google Scholar]
  105. Stinziano JR, Way DA, Bauerle WL.. 2018. Improving models of photosynthetic thermal acclimation: which parameters are most important and how many should be modified? Global Change Biology 24, 1580–1598. [DOI] [PubMed] [Google Scholar]
  106. Thioulouse J, Dray S, Dufour A-B, Siberchicot A, Jombart T, Pavoine, S.. 2018. Multivariate analysis of ecological data with ade4. New York: Springer. [Google Scholar]
  107. Tsugawa H, Ikeda K, Takahashi M, et al. 2020. A lipidome atlas in MS-DIAL 4. Nature Biotechnology 38, 1159–1163. [DOI] [PubMed] [Google Scholar]
  108. Turnbull MH, Murthy R, Griffin KL.. 2002. The relative impacts of daytime and night-time warming on photosynthetic capacity in Populus deltoides. Plant, Cell & Environment 25, 1729–1737. [Google Scholar]
  109. Urban J, Ingwers M, Mcguire MA, Teskey RO.. 2017a. Stomatal conductance increases with rising temperature. Plant Signaling & Behavior 12, e1356534. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Urban J, Ingwers M, Mcguire MA, Teskey RO.. 2017b. Increase in leaf temperature opens stomata and decouples net photosynthesis from stomatal conductance in Pinus taeda and Populus deltoides × nigra. Journal of Experimental Botany 68, 1757–1767. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Verhoef A, Egea G.. 2014. Modeling plant transpiration under limited soilwater: comparison of different plant and soil hydraulic parameterizations and preliminary implications for their use in land surface models. Agricultural and Forest Meteorology 191, 22–32. [Google Scholar]
  112. von Caemmerer S. 2000. Biochemical models of leaf photosynthesis, Collingwood, Victoria: CSIRO Publishing. [Google Scholar]
  113. von Caemmerer S, Evans JR.. 2015. Temperature responses of mesophyll conductance differ greatly between species. Plant, Cell & Environment 38, 629–637. [DOI] [PubMed] [Google Scholar]
  114. Wang C, Caragea D, Kodadinne Narayana N, Hein NT, Bheemanahalli R, Somayanda IM, Jagadish SVK.. 2022. Deep learning based high-throughput phenotyping of chalkiness in rice exposed to high night temperature. Plant Methods 18, 9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Way DA, Yamori W.. 2014. Thermal acclimation of photosynthesis: on the importance of adjusting our definitions and accounting for thermal acclimation of respiration. Photosynthesis Research 119, 89–100. [DOI] [PubMed] [Google Scholar]
  116. Wickham H. 2016. ggplot2: elegrant graphics for data analysis. New York: Springer. [Google Scholar]
  117. Wickham H. 2021. tidyr: tidy messy data. R package version. 1.1.3. https://cran.r-project.org/web/packages/tidyr/index.html [Google Scholar]
  118. Wickham H, Averick M, Bryan J, Chang W, Mcgowan LDA, François R, Grolemund G, Hayes A, Henry L, Hester J.. 2019. Welcome to the Tidyverse. Journal of Open Source Software 4, 1686. [Google Scholar]
  119. Wickham H., François R, Henry L, Müller K.. 2021. dplyr: a grammar of data manipulation. R Package Version 1.0.6. https://cran.r-project.org/web/packages/dplyr/index.html [Google Scholar]
  120. Yamasaki T, Yamakawa T, Yamane Y, Koike H, Satoh K, Katoh S.. 2002. Temperature acclimation of photosynthesis and related changes in photosystem II electron transport in winter wheat. Plant Physiology 128, 1087–1097. [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Yamori W, Noguchi K, Terashima I.. 2005. Temperature acclimation of photosynthesis in spinach leaves: analyses of photosynthetic components and temperature dependencies of photosynthetic partial reactions. Plant, Cell & Environment 28, 536–547. [Google Scholar]
  122. Yamori W, Suzuki K, Noguchi KO, Nakai M, Terashima I.. 2006. Effects of Rubisco kinetics and Rubisco activation state on the temperature dependence of the photosynthetic rate in spinach leaves from contrasting growth temperatures. Plant, Cell & Environment 29, 1659–1670. [DOI] [PubMed] [Google Scholar]
  123. Zadoks JC, Chang TT, Konzak CF.. 1974. A decimal code for the growth stages of cereals. Weed Research 14, 415–421. [Google Scholar]
  124. Zaicou-Kunesch C, Trainor G, Shackley B, Curry J, Nicol D, Shankar M, Huberli D.Dhammu H.. 2017. Wheat variety fact sheets for Western Australia. Western Australia: Department of Agriculture and Food. [Google Scholar]
  125. Zhu L, Bloomfield KJ, Hocart CH, Egerton JJG, O’Sullivan OS, Penillard A, Weerasinghe LK, Atkin OK.. 2018. Plasticity of photosynthetic heat tolerance in plants adapted to thermally contrasting biomes. Plant, Cell & Environment 41, 1251–1262. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

erad437_suppl_Supplementary_Figures_S1-S2_Tables_S1
erad437_suppl_Supplementary_Tables_S2

Data Availability Statement

All primary data to support the findings of this study are openly available in the Dryad Digital Repository at doi: 10.5061/dryad.fqz612jx7 (Coast et al., 2023).


Articles from Journal of Experimental Botany are provided here courtesy of Oxford University Press

RESOURCES